01.08.2021 Views

Variational Principles in Classical Mechanics - Revised Second Edition, 2019

Variational Principles in Classical Mechanics - Revised Second Edition, 2019

Variational Principles in Classical Mechanics - Revised Second Edition, 2019

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

VARIATIONAL PRINCIPLES<br />

IN<br />

CLASSICAL MECHANICS<br />

REVISED SECOND EDITION<br />

Douglas Cl<strong>in</strong>e<br />

University of Rochester<br />

15 January <strong>2019</strong>


ii<br />

c°<strong>2019</strong>, 2017 by Douglas Cl<strong>in</strong>e<br />

ISBN: 978-0-9988372-8-4 e-book (Adobe PDF)<br />

ISBN: 978-0-9988372-2-2 e-book (K<strong>in</strong>dle)<br />

ISBN: 978-0-9988372-9-1 pr<strong>in</strong>t (Paperback)<br />

<strong>Variational</strong> <strong>Pr<strong>in</strong>ciples</strong> <strong>in</strong> <strong>Classical</strong> <strong>Mechanics</strong>, <strong>Revised</strong> 2 edition<br />

Contributors<br />

Author: Douglas Cl<strong>in</strong>e<br />

Illustrator: Meghan Sarkis<br />

Published by University of Rochester River Campus Libraries<br />

University of Rochester<br />

Rochester, NY 14627<br />

<strong>Variational</strong> <strong>Pr<strong>in</strong>ciples</strong> <strong>in</strong> <strong>Classical</strong> <strong>Mechanics</strong>, <strong>Revised</strong> 2 edition by Douglas Cl<strong>in</strong>e is licensed under a<br />

Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License (CC BY-NC-SA 4.0),<br />

except where otherwise noted.<br />

You are free to:<br />

• Share — copy or redistribute the material <strong>in</strong> any medium or format.<br />

• Adapt — remix, transform, and build upon the material.<br />

Under the follow<strong>in</strong>g terms:<br />

• Attribution — You must give appropriate credit, provide a l<strong>in</strong>k to the license, and <strong>in</strong>dicate if changes<br />

were made. You must do so <strong>in</strong> any reasonable manner, but not <strong>in</strong> any way that suggests the licensor<br />

endorsesyouoryouruse.<br />

• NonCommercial — You may not use the material for commercial purposes.<br />

• ShareAlike — If you remix, transform, or build upon the material, you must distribute your contributions<br />

under the same license as the orig<strong>in</strong>al.<br />

• No additional restrictions — You may not apply legal terms or technological measures that legally<br />

restrict others from do<strong>in</strong>g anyth<strong>in</strong>g the license permits.<br />

Thelicensorcannotrevokethesefreedomsaslongasyoufollowthelicenseterms.<br />

Version 2.1


Contents<br />

Contents<br />

Preface<br />

Prologue<br />

iii<br />

xvii<br />

xix<br />

1 A brief history of classical mechanics 1<br />

1.1 Introduction.............................................. 1<br />

1.2 Greekantiquity............................................ 1<br />

1.3 MiddleAges.............................................. 2<br />

1.4 AgeofEnlightenment ........................................ 2<br />

1.5 <strong>Variational</strong>methods<strong>in</strong>physics ................................... 5<br />

1.6 The 20 centuryrevolution<strong>in</strong>physics............................... 7<br />

2 ReviewofNewtonianmechanics 9<br />

2.1 Introduction.............................................. 9<br />

2.2 Newton’sLawsofmotion ...................................... 9<br />

2.3 Inertialframesofreference...................................... 10<br />

2.4 First-order<strong>in</strong>tegrals<strong>in</strong>Newtonianmechanics ........................... 11<br />

2.4.1 L<strong>in</strong>earMomentum ...................................... 11<br />

2.4.2 Angularmomentum ..................................... 11<br />

2.4.3 K<strong>in</strong>eticenergy ........................................ 12<br />

2.5 Conservationlaws<strong>in</strong>classicalmechanics.............................. 12<br />

2.6 Motion of f<strong>in</strong>ite-sizedandmany-bodysystems........................... 12<br />

2.7 Centerofmassofamany-bodysystem............................... 13<br />

2.8 Totall<strong>in</strong>earmomentumofamany-bodysystem.......................... 14<br />

2.8.1 Center-of-massdecomposition................................ 14<br />

2.8.2 Equationsofmotion ..................................... 14<br />

2.9 Angularmomentumofamany-bodysystem............................ 16<br />

2.9.1 Center-of-massdecomposition................................ 16<br />

2.9.2 Equationsofmotion ..................................... 16<br />

2.10Workandk<strong>in</strong>eticenergyforamany-bodysystem......................... 18<br />

2.10.1 Center-of-massk<strong>in</strong>eticenergy................................ 18<br />

2.10.2 Conservativeforcesandpotentialenergy.......................... 18<br />

2.10.3 Totalmechanicalenergy................................... 19<br />

2.10.4 Totalmechanicalenergyforconservativesystems..................... 20<br />

2.11VirialTheorem ............................................ 22<br />

2.12ApplicationsofNewton’sequationsofmotion........................... 24<br />

2.12.1 Constantforceproblems................................... 24<br />

2.12.2 L<strong>in</strong>earRestor<strong>in</strong>gForce.................................... 25<br />

2.12.3 Position-dependentconservativeforces........................... 25<br />

2.12.4 Constra<strong>in</strong>edmotion ..................................... 27<br />

2.12.5 VelocityDependentForces.................................. 28<br />

2.12.6 SystemswithVariableMass................................. 29<br />

2.12.7 Rigid-body rotation about a body-fixedrotationaxis................... 31<br />

iii


iv<br />

CONTENTS<br />

2.12.8 Timedependentforces.................................... 34<br />

2.13Solutionofmany-bodyequationsofmotion ............................ 37<br />

2.13.1 Analyticsolution....................................... 37<br />

2.13.2 Successiveapproximation .................................. 37<br />

2.13.3 Perturbationmethod..................................... 37<br />

2.14Newton’sLawofGravitation .................................... 38<br />

2.14.1 Gravitationaland<strong>in</strong>ertialmass............................... 38<br />

2.14.2 Gravitational potential energy .............................. 39<br />

2.14.3 Gravitational potential .................................. 40<br />

2.14.4 Potentialtheory ....................................... 41<br />

2.14.5 Curl of the gravitational field................................ 41<br />

2.14.6 Gauss’sLawforGravitation................................. 43<br />

2.14.7 CondensedformsofNewton’sLawofGravitation..................... 44<br />

2.15Summary ............................................... 46<br />

Problems .................................................. 48<br />

3 L<strong>in</strong>ear oscillators 51<br />

3.1 Introduction.............................................. 51<br />

3.2 L<strong>in</strong>earrestor<strong>in</strong>gforces ........................................ 51<br />

3.3 L<strong>in</strong>earityandsuperposition ..................................... 52<br />

3.4 Geometricalrepresentationsofdynamicalmotion......................... 53<br />

3.4.1 Configuration space ( ) ................................ 53<br />

3.4.2 State space, ( ˙ ) ..................................... 54<br />

3.4.3 Phase space, ( ) .................................... 54<br />

3.4.4 Planependulum ....................................... 55<br />

3.5 L<strong>in</strong>early-damped free l<strong>in</strong>ear oscillator ................................ 56<br />

3.5.1 Generalsolution ....................................... 56<br />

3.5.2 Energydissipation ...................................... 59<br />

3.6 S<strong>in</strong>usoidally-drive, l<strong>in</strong>early-damped, l<strong>in</strong>ear oscillator . . . .................... 60<br />

3.6.1 Transient response of a driven oscillator .......................... 60<br />

3.6.2 Steady state response of a driven oscillator ........................ 61<br />

3.6.3 Completesolutionofthedrivenoscillator ......................... 62<br />

3.6.4 Resonance........................................... 63<br />

3.6.5 Energyabsorption ...................................... 63<br />

3.7 Waveequation ............................................ 66<br />

3.8 Travell<strong>in</strong>gandstand<strong>in</strong>gwavesolutionsofthewaveequation................... 67<br />

3.9 Waveformanalysis .......................................... 68<br />

3.9.1 Harmonicdecomposition................................... 68<br />

3.9.2 The free l<strong>in</strong>early-damped l<strong>in</strong>ear oscillator ......................... 68<br />

3.9.3 Dampedl<strong>in</strong>earoscillatorsubjecttoanarbitraryperiodicforce ............. 69<br />

3.10Signalprocess<strong>in</strong>g ........................................... 70<br />

3.11 Wave propagation .......................................... 71<br />

3.11.1 Phase,group,andsignalvelocitiesofwavepackets .................... 72<br />

3.11.2 Fouriertransformofwavepackets ............................. 77<br />

3.11.3 Wave-packetUncerta<strong>in</strong>tyPr<strong>in</strong>ciple............................. 78<br />

3.12Summary ............................................... 80<br />

Problems .................................................. 83<br />

4 Nonl<strong>in</strong>ear systems and chaos 85<br />

4.1 Introduction.............................................. 85<br />

4.2 Weaknonl<strong>in</strong>earity .......................................... 86<br />

4.3 Bifurcation,andpo<strong>in</strong>tattractors .................................. 88<br />

4.4 Limitcycles.............................................. 89<br />

4.4.1 Po<strong>in</strong>caré-Bendixsontheorem ................................ 89<br />

4.4.2 vanderPoldampedharmonicoscillator:.......................... 90


CONTENTS<br />

v<br />

4.5 Harmonically-driven,l<strong>in</strong>early-damped,planependulum...................... 93<br />

4.5.1 Closetol<strong>in</strong>earity....................................... 93<br />

4.5.2 Weaknonl<strong>in</strong>earity ...................................... 95<br />

4.5.3 Onsetofcomplication .................................... 96<br />

4.5.4 Perioddoubl<strong>in</strong>gandbifurcation............................... 96<br />

4.5.5 Roll<strong>in</strong>g motion ........................................ 96<br />

4.5.6 Onsetofchaos ........................................ 97<br />

4.6 Differentiationbetweenorderedandchaoticmotion........................ 98<br />

4.6.1 Lyapunovexponent ..................................... 98<br />

4.6.2 Bifurcationdiagram ..................................... 99<br />

4.6.3 Po<strong>in</strong>caréSection .......................................100<br />

4.7 Wave propagation for non-l<strong>in</strong>ear systems . . . ...........................101<br />

4.7.1 Phase,group,andsignalvelocities .............................101<br />

4.7.2 Soliton wave propagation ..................................103<br />

4.8 Summary ...............................................104<br />

Problems ..................................................106<br />

5 Calculus of variations 107<br />

5.1 Introduction..............................................107<br />

5.2 Euler’s differentialequation .....................................108<br />

5.3 ApplicationsofEuler’sequation...................................110<br />

5.4 Selectionofthe<strong>in</strong>dependentvariable................................113<br />

5.5 Functions with several <strong>in</strong>dependent variables () ........................115<br />

5.6 Euler’s<strong>in</strong>tegralequation.......................................117<br />

5.7 Constra<strong>in</strong>edvariationalsystems...................................118<br />

5.7.1 Holonomicconstra<strong>in</strong>ts....................................118<br />

5.7.2 Geometric(algebraic)equationsofconstra<strong>in</strong>t .......................118<br />

5.7.3 K<strong>in</strong>ematic (differential)equationsofconstra<strong>in</strong>t ......................118<br />

5.7.4 Isoperimetric(<strong>in</strong>tegral)equationsofconstra<strong>in</strong>t ......................119<br />

5.7.5 Propertiesoftheconstra<strong>in</strong>tequations ...........................119<br />

5.7.6 Treatmentofconstra<strong>in</strong>tforces<strong>in</strong>variationalcalculus...................120<br />

5.8 Generalizedcoord<strong>in</strong>ates<strong>in</strong>variationalcalculus ..........................121<br />

5.9 Lagrangemultipliersforholonomicconstra<strong>in</strong>ts ..........................122<br />

5.9.1 Algebraicequationsofconstra<strong>in</strong>t..............................122<br />

5.9.2 Integralequationsofconstra<strong>in</strong>t...............................124<br />

5.10Geodesic................................................126<br />

5.11<strong>Variational</strong>approachtoclassicalmechanics ............................127<br />

5.12Summary ...............................................128<br />

Problems ..................................................129<br />

6 Lagrangian dynamics 131<br />

6.1 Introduction..............................................131<br />

6.2 Newtonian plausibility argument for Lagrangian mechanics ...................132<br />

6.3 Lagrangeequationsfromd’Alembert’sPr<strong>in</strong>ciple..........................134<br />

6.3.1 d’Alembert’sPr<strong>in</strong>cipleofVirtualWork ..........................134<br />

6.3.2 Transformationtogeneralizedcoord<strong>in</strong>ates.........................135<br />

6.3.3 Lagrangian ..........................................136<br />

6.4 LagrangeequationsfromHamilton’sActionPr<strong>in</strong>ciple ......................137<br />

6.5 Constra<strong>in</strong>edsystems .........................................138<br />

6.5.1 Choiceofgeneralizedcoord<strong>in</strong>ates..............................138<br />

6.5.2 M<strong>in</strong>imalsetofgeneralizedcoord<strong>in</strong>ates...........................138<br />

6.5.3 Lagrangemultipliersapproach ...............................138<br />

6.5.4 Generalizedforcesapproach.................................140<br />

6.6 Apply<strong>in</strong>gtheEuler-Lagrangeequationstoclassicalmechanics..................140<br />

6.7 Applicationstounconstra<strong>in</strong>edsystems...............................142


vi<br />

CONTENTS<br />

6.8 Applicationstosystems<strong>in</strong>volv<strong>in</strong>gholonomicconstra<strong>in</strong>ts .....................144<br />

6.9 Applications<strong>in</strong>volv<strong>in</strong>gnon-holonomicconstra<strong>in</strong>ts.........................157<br />

6.10Velocity-dependentLorentzforce ..................................164<br />

6.11Time-dependentforces........................................165<br />

6.12Impulsiveforces............................................166<br />

6.13TheLagrangianversustheNewtonianapproachtoclassicalmechanics.............168<br />

6.14Summary ...............................................169<br />

Problems ..................................................172<br />

7 Symmetries, Invariance and the Hamiltonian 175<br />

7.1 Introduction..............................................175<br />

7.2 Generalizedmomentum .......................................175<br />

7.3 InvarianttransformationsandNoether’sTheorem.........................177<br />

7.4 Rotational<strong>in</strong>varianceandconservationofangularmomentum..................179<br />

7.5 Cycliccoord<strong>in</strong>ates ..........................................180<br />

7.6 K<strong>in</strong>eticenergy<strong>in</strong>generalizedcoord<strong>in</strong>ates .............................181<br />

7.7 GeneralizedenergyandtheHamiltonianfunction.........................182<br />

7.8 Generalizedenergytheorem.....................................183<br />

7.9 Generalizedenergyandtotalenergy ................................183<br />

7.10Hamiltonian<strong>in</strong>variance........................................184<br />

7.11Hamiltonianforcycliccoord<strong>in</strong>ates .................................189<br />

7.12Symmetriesand<strong>in</strong>variance .....................................189<br />

7.13Hamiltonian<strong>in</strong>classicalmechanics .................................189<br />

7.14Summary ...............................................190<br />

Problems ..................................................192<br />

8 Hamiltonian mechanics 195<br />

8.1 Introduction..............................................195<br />

8.2 LegendreTransformationbetweenLagrangianandHamiltonianmechanics...........196<br />

8.3 Hamilton’sequationsofmotion...................................197<br />

8.3.1 Canonicalequationsofmotion ...............................198<br />

8.4 Hamiltonian <strong>in</strong> differentcoord<strong>in</strong>atesystems............................199<br />

8.4.1 Cyl<strong>in</strong>drical coord<strong>in</strong>ates ................................199<br />

8.4.2 Spherical coord<strong>in</strong>ates, .................................200<br />

8.5 ApplicationsofHamiltonianDynamics...............................201<br />

8.6 Routhianreduction..........................................206<br />

8.6.1 R -RouthianisaHamiltonianforthecyclicvariables................207<br />

8.6.2 R -RouthianisaHamiltonianforthenon-cyclicvariables ...........208<br />

8.7 Variable-masssystems ........................................212<br />

8.7.1 Rocketpropulsion:......................................212<br />

8.7.2 Mov<strong>in</strong>gcha<strong>in</strong>s: ........................................213<br />

8.8 Summary ...............................................215<br />

Problems ..................................................217<br />

9 Hamilton’s Action Pr<strong>in</strong>ciple 221<br />

9.1 Introduction..............................................221<br />

9.2 Hamilton’s Pr<strong>in</strong>ciple of Stationary Action<br />

9.2.1 Stationary-actionpr<strong>in</strong>ciple<strong>in</strong>Lagrangianmechanics ...................222<br />

9.2.2 Stationary-actionpr<strong>in</strong>ciple<strong>in</strong>Hamiltonianmechanics ..................223<br />

9.2.3 Abbreviatedaction......................................224<br />

9.2.4 Hamilton’s Pr<strong>in</strong>ciple applied us<strong>in</strong>g <strong>in</strong>itial boundary conditions . ............225<br />

9.3 Lagrangian ..............................................228<br />

9.3.1 StandardLagrangian.....................................228<br />

9.3.2 Gauge<strong>in</strong>varianceofthestandardLagrangian .......................228<br />

9.3.3 Non-standardLagrangians..................................230<br />

9.3.4 Inversevariationalcalculus .................................230


CONTENTS<br />

vii<br />

9.4 ApplicationofHamilton’sActionPr<strong>in</strong>cipletomechanics.....................231<br />

9.5 Summary ...............................................232<br />

10 Nonconservative systems 235<br />

10.1Introduction..............................................235<br />

10.2Orig<strong>in</strong>sofnonconservativemotion .................................235<br />

10.3Algebraicmechanicsfornonconservativesystems .........................236<br />

10.4Rayleigh’sdissipationfunction ...................................236<br />

10.4.1 Generalizeddissipativeforcesforl<strong>in</strong>earvelocitydependence...............237<br />

10.4.2 Generalizeddissipativeforcesfornonl<strong>in</strong>earvelocitydependence.............238<br />

10.4.3 Lagrangeequationsofmotion................................238<br />

10.4.4 Hamiltonianmechanics ...................................238<br />

10.5DissipativeLagrangians .......................................241<br />

10.6Summary ...............................................243<br />

11 Conservative two-body central forces 245<br />

11.1Introduction..............................................245<br />

11.2Equivalentone-bodyrepresentationfortwo-bodymotion.....................246<br />

11.3 Angular momentum L ........................................248<br />

11.4Equationsofmotion .........................................249<br />

11.5 Differentialorbitequation:......................................250<br />

11.6Hamiltonian..............................................251<br />

11.7Generalfeaturesoftheorbitsolutions ...............................252<br />

11.8Inverse-square,two-body,centralforce...............................253<br />

11.8.1 Boundorbits .........................................254<br />

11.8.2 Kepler’slawsforboundplanetarymotion .........................255<br />

11.8.3 Unboundorbits........................................256<br />

11.8.4 Eccentricity vector . . . ...................................257<br />

11.9Isotropic,l<strong>in</strong>ear,two-body,centralforce ..............................259<br />

11.9.1 Polarcoord<strong>in</strong>ates.......................................260<br />

11.9.2 Cartesiancoord<strong>in</strong>ates ....................................261<br />

11.9.3 Symmetry tensor A 0 .....................................262<br />

11.10Closed-orbitstability.........................................263<br />

11.11Thethree-bodyproblem.......................................268<br />

11.12Two-bodyscatter<strong>in</strong>g.........................................269<br />

11.12.1Totaltwo-bodyscatter<strong>in</strong>gcrosssection ..........................269<br />

11.12.2 Differentialtwo-bodyscatter<strong>in</strong>gcrosssection .......................270<br />

11.12.3Impactparameterdependenceonscatter<strong>in</strong>gangle ....................270<br />

11.12.4Rutherfordscatter<strong>in</strong>g ....................................272<br />

11.13Two-bodyk<strong>in</strong>ematics.........................................274<br />

11.14Summary ...............................................280<br />

Problems ..................................................282<br />

12 Non-<strong>in</strong>ertial reference frames 285<br />

12.1Introduction..............................................285<br />

12.2 Translational acceleration of a reference frame ...........................285<br />

12.3Rotat<strong>in</strong>greferenceframe.......................................286<br />

12.3.1 Spatialtimederivatives<strong>in</strong>arotat<strong>in</strong>g,non-translat<strong>in</strong>g,referenceframe .........286<br />

12.3.2 Generalvector<strong>in</strong>arotat<strong>in</strong>g,non-translat<strong>in</strong>g,referenceframe ..............287<br />

12.4 Reference frame undergo<strong>in</strong>g rotation plus translation . . . ....................288<br />

12.5Newton’slawofmotion<strong>in</strong>anon-<strong>in</strong>ertialframe ..........................288<br />

12.6Lagrangianmechanics<strong>in</strong>anon-<strong>in</strong>ertialframe ...........................289<br />

12.7Centrifugalforce ...........................................290<br />

12.8Coriolisforce .............................................291<br />

12.9Routhianreductionforrotat<strong>in</strong>gsystems ..............................295<br />

12.10EffectivegravitationalforcenearthesurfaceoftheEarth ....................298


viii<br />

CONTENTS<br />

12.11Freemotionontheearth.......................................300<br />

12.12Weathersystems ...........................................302<br />

12.12.1Low-pressuresystems: ....................................302<br />

12.12.2High-pressuresystems:....................................304<br />

12.13Foucaultpendulum..........................................304<br />

12.14Summary ...............................................306<br />

Problems ..................................................307<br />

13 Rigid-body rotation 309<br />

13.1Introduction..............................................309<br />

13.2Rigid-bodycoord<strong>in</strong>ates........................................310<br />

13.3 Rigid-body rotation about a body-fixedpo<strong>in</strong>t...........................310<br />

13.4Inertiatensor .............................................312<br />

13.5Matrixandtensorformulationsofrigid-bodyrotation ......................313<br />

13.6Pr<strong>in</strong>cipalaxissystem.........................................313<br />

13.7 Diagonalize the <strong>in</strong>ertia tensor . ...................................314<br />

13.8Parallel-axistheorem.........................................315<br />

13.9Perpendicular-axistheoremforplanelam<strong>in</strong>ae ...........................318<br />

13.10Generalpropertiesofthe<strong>in</strong>ertiatensor...............................319<br />

13.10.1Inertialequivalence......................................319<br />

13.10.2 Orthogonality of pr<strong>in</strong>cipal axes ...............................320<br />

13.11Angular momentum L and angular velocity ω vectors ......................321<br />

13.12K<strong>in</strong>eticenergyofrotat<strong>in</strong>grigidbody................................323<br />

13.13Eulerangles..............................................325<br />

13.14Angular velocity ω ..........................................327<br />

13.15K<strong>in</strong>eticenergy<strong>in</strong>termsofEulerangularvelocities ........................328<br />

13.16Rotational<strong>in</strong>variants.........................................329<br />

13.17Euler’sequationsofmotionforrigid-bodyrotation ........................330<br />

13.18Lagrangeequationsofmotionforrigid-bodyrotation.......................331<br />

13.19Hamiltonianequationsofmotionforrigid-bodyrotation.....................333<br />

13.20Torque-freerotationofan<strong>in</strong>ertially-symmetricrigidrotor ....................333<br />

13.20.1Euler’sequationsofmotion:.................................333<br />

13.20.2Lagrangeequationsofmotion: ...............................337<br />

13.21Torque-freerotationofanasymmetricrigidrotor.........................339<br />

13.22Stability of torque-free rotation of an asymmetric body . . ....................340<br />

13.23Symmetric rigid rotor subject to torque about a fixedpo<strong>in</strong>t ...................343<br />

13.24Theroll<strong>in</strong>gwheel...........................................347<br />

13.25Dynamicbalanc<strong>in</strong>gofwheels ....................................350<br />

13.26Rotationofdeformablebodies....................................351<br />

13.27Summary ...............................................352<br />

Problems ..................................................354<br />

14 Coupled l<strong>in</strong>ear oscillators 357<br />

14.1Introduction..............................................357<br />

14.2 Two coupled l<strong>in</strong>ear oscillators . ...................................357<br />

14.3Normalmodes ............................................359<br />

14.4 Center of mass oscillations . . . ...................................360<br />

14.5Weakcoupl<strong>in</strong>g ............................................361<br />

14.6Generalanalytictheoryforcoupledl<strong>in</strong>earoscillators .......................363<br />

14.6.1 K<strong>in</strong>etic energy tensor T ...................................363<br />

14.6.2 Potential energy tensor V ..................................364<br />

14.6.3 Equationsofmotion .....................................365<br />

14.6.4 Superposition.........................................366<br />

14.6.5 Eigenfunctionorthonormality................................366<br />

14.6.6 Normalcoord<strong>in</strong>ates .....................................367


CONTENTS<br />

ix<br />

14.7 Two-body coupled oscillator systems ................................368<br />

14.8 Three-body coupled l<strong>in</strong>ear oscillator systems . ...........................374<br />

14.9Molecularcoupledoscillatorsystems ................................379<br />

14.10DiscreteLatticeCha<strong>in</strong>........................................382<br />

14.10.1Longitud<strong>in</strong>almotion.....................................382<br />

14.10.2Transversemotion ......................................382<br />

14.10.3Normalmodes........................................383<br />

14.10.4 Travell<strong>in</strong>g waves .......................................386<br />

14.10.5Dispersion...........................................386<br />

14.10.6Complexwavenumber ....................................387<br />

14.11Damped coupled l<strong>in</strong>ear oscillators ..................................388<br />

14.12Collective synchronization of coupled oscillators ..........................389<br />

14.13Summary ...............................................392<br />

Problems ..................................................393<br />

15 Advanced Hamiltonian mechanics 395<br />

15.1Introduction..............................................395<br />

15.2PoissonbracketrepresentationofHamiltonianmechanics ....................397<br />

15.2.1 PoissonBrackets .......................................397<br />

15.2.2 FundamentalPoissonbrackets: ...............................397<br />

15.2.3 Poissonbracket<strong>in</strong>variancetocanonicaltransformations .................398<br />

15.2.4 CorrespondenceofthecommutatorandthePoissonBracket...............399<br />

15.2.5 Observables<strong>in</strong>Hamiltonianmechanics...........................400<br />

15.2.6 Hamilton’sequationsofmotion...............................403<br />

15.2.7 Liouville’s Theorem . . ...................................407<br />

15.3Canonicaltransformations<strong>in</strong>Hamiltonianmechanics.......................409<br />

15.3.1 Generat<strong>in</strong>gfunctions.....................................410<br />

15.3.2 Applicationsofcanonicaltransformations .........................412<br />

15.4Hamilton-Jacobitheory .......................................414<br />

15.4.1 Time-dependentHamiltonian................................414<br />

15.4.2 Time-<strong>in</strong>dependentHamiltonian...............................416<br />

15.4.3 Separationofvariables....................................417<br />

15.4.4 Visual representation of the action function . ......................424<br />

15.4.5 AdvantagesofHamilton-Jacobitheory...........................424<br />

15.5Action-anglevariables ........................................425<br />

15.5.1 Canonicaltransformation ..................................425<br />

15.5.2 Adiabatic<strong>in</strong>varianceoftheactionvariables ........................428<br />

15.6Canonicalperturbationtheory ...................................430<br />

15.7Symplecticrepresentation ......................................432<br />

15.8ComparisonoftheLagrangianandHamiltonianformulations ..................432<br />

15.9Summary ...............................................434<br />

Problems ..................................................437<br />

16 Analytical formulations for cont<strong>in</strong>uous systems 439<br />

16.1Introduction..............................................439<br />

16.2Thecont<strong>in</strong>uousuniforml<strong>in</strong>earcha<strong>in</strong> ................................439<br />

16.3TheLagrangiandensityformulationforcont<strong>in</strong>uoussystems ...................440<br />

16.3.1 Onespatialdimension....................................440<br />

16.3.2 Threespatialdimensions ..................................441<br />

16.4TheHamiltoniandensityformulationforcont<strong>in</strong>uoussystems ..................442<br />

16.5L<strong>in</strong>earelasticsolids..........................................443<br />

16.5.1 Stresstensor .........................................444<br />

16.5.2 Stra<strong>in</strong>tensor .........................................444<br />

16.5.3 Moduliofelasticity......................................445<br />

16.5.4 Equationsofmotion<strong>in</strong>auniformelasticmedia......................446


x<br />

CONTENTS<br />

16.6 Electromagnetic fieldtheory.....................................447<br />

16.6.1 Maxwellstresstensor ....................................447<br />

16.6.2 Momentum <strong>in</strong> the electromagnetic field ..........................448<br />

16.7 Ideal fluiddynamics .........................................449<br />

16.7.1 Cont<strong>in</strong>uityequation .....................................449<br />

16.7.2 Euler’shydrodynamicequation...............................449<br />

16.7.3 Irrotational flow and Bernoulli’s equation .........................450<br />

16.7.4 Gas flow............................................450<br />

16.8 Viscous fluiddynamics........................................452<br />

16.8.1 Navier-Stokesequation....................................452<br />

16.8.2 Reynoldsnumber.......................................453<br />

16.8.3 Lam<strong>in</strong>ar and turbulent fluid flow..............................453<br />

16.9Summaryandimplications .....................................455<br />

17 Relativistic mechanics 457<br />

17.1Introduction..............................................457<br />

17.2GalileanInvariance..........................................457<br />

17.3SpecialTheoryofRelativity.....................................459<br />

17.3.1 E<strong>in</strong>ste<strong>in</strong>Postulates......................................459<br />

17.3.2 Lorentztransformation ...................................459<br />

17.3.3 TimeDilation: ........................................460<br />

17.3.4 LengthContraction .....................................461<br />

17.3.5 Simultaneity .........................................461<br />

17.4Relativistick<strong>in</strong>ematics........................................464<br />

17.4.1 Velocitytransformations...................................464<br />

17.4.2 Momentum ..........................................464<br />

17.4.3 Centerofmomentumcoord<strong>in</strong>atesystem..........................465<br />

17.4.4 Force .............................................465<br />

17.4.5 Energy.............................................465<br />

17.5Geometryofspace-time .......................................467<br />

17.5.1 Four-dimensionalspace-time ................................467<br />

17.5.2 Four-vectorscalarproducts .................................468<br />

17.5.3 M<strong>in</strong>kowskispace-time ....................................469<br />

17.5.4 Momentum-energyfourvector ...............................470<br />

17.6Lorentz-<strong>in</strong>variantformulationofLagrangianmechanics......................471<br />

17.6.1 Parametricformulation ...................................471<br />

17.6.2 ExtendedLagrangian ....................................471<br />

17.6.3 Extendedgeneralizedmomenta...............................473<br />

17.6.4 ExtendedLagrangeequationsofmotion ..........................473<br />

17.7Lorentz-<strong>in</strong>variantformulationsofHamiltonianmechanics.....................476<br />

17.7.1 Extendedcanonicalformalism................................476<br />

17.7.2 ExtendedPoissonBracketrepresentation .........................478<br />

17.7.3 ExtendedcanonicaltransformationandHamilton-Jacobitheory.............478<br />

17.7.4 ValidityoftheextendedHamilton-Lagrangeformalism..................478<br />

17.8TheGeneralTheoryofRelativity..................................480<br />

17.8.1 Thefundamentalconcepts..................................480<br />

17.8.2 E<strong>in</strong>ste<strong>in</strong>’spostulatesfortheGeneralTheoryofRelativity ................481<br />

17.8.3 Experimentalevidence<strong>in</strong>supportoftheGeneralTheoryofRelativity .........481<br />

17.9Implicationsofrelativistictheorytoclassicalmechanics .....................482<br />

17.10Summary ...............................................483<br />

Problems ..................................................484<br />

18 The transition to quantum physics 485<br />

18.1Introduction..............................................485<br />

18.2Briefsummaryoftheorig<strong>in</strong>sofquantumtheory .........................485


CONTENTS<br />

xi<br />

18.2.1 Bohrmodeloftheatom...................................487<br />

18.2.2 Quantization .........................................487<br />

18.2.3 Wave-particleduality ....................................488<br />

18.3Hamiltonian<strong>in</strong>quantumtheory...................................489<br />

18.3.1 Heisenberg’smatrix-mechanicsrepresentation.......................489<br />

18.3.2 Schröd<strong>in</strong>ger’swave-mechanicsrepresentation .......................491<br />

18.4Lagrangianrepresentation<strong>in</strong>quantumtheory...........................492<br />

18.5CorrespondencePr<strong>in</strong>ciple ......................................493<br />

18.6Summary ...............................................494<br />

19 Epilogue 495<br />

Appendices<br />

A Matrix algebra 497<br />

A.1 Mathematicalmethodsformechanics................................497<br />

A.2 Matrices................................................497<br />

A.3 Determ<strong>in</strong>ants .............................................501<br />

A.4 Reduction of a matrix to diagonal form . . . ...........................503<br />

B Vector algebra 505<br />

B.1 L<strong>in</strong>earoperations...........................................505<br />

B.2 Scalarproduct ............................................505<br />

B.3 Vectorproduct ............................................506<br />

B.4 Tripleproducts............................................507<br />

C Orthogonal coord<strong>in</strong>ate systems 509<br />

C.1 Cartesian coord<strong>in</strong>ates ( ) ....................................509<br />

C.2 Curvil<strong>in</strong>ear coord<strong>in</strong>ate systems ...................................509<br />

C.2.1 Two-dimensional polar coord<strong>in</strong>ates ( ) ..........................510<br />

C.2.2 Cyl<strong>in</strong>drical Coord<strong>in</strong>ates ( ) ..............................512<br />

C.2.3 Spherical Coord<strong>in</strong>ates ( ) ................................512<br />

C.3 Frenet-Serretcoord<strong>in</strong>ates ......................................513<br />

D Coord<strong>in</strong>ate transformations 515<br />

D.1 Translationaltransformations....................................515<br />

D.2 Rotationaltransformations .....................................515<br />

D.2.1 Rotationmatrix .......................................515<br />

D.2.2 F<strong>in</strong>iterotations........................................518<br />

D.2.3 Inf<strong>in</strong>itessimalrotations....................................519<br />

D.2.4 Properandimproperrotations ...............................519<br />

D.3 Spatial<strong>in</strong>versiontransformation...................................520<br />

D.4 Timereversaltransformation ....................................521<br />

E Tensor algebra 523<br />

E.1 Tensors ................................................523<br />

E.2 Tensorproducts............................................524<br />

E.2.1 Tensorouterproduct.....................................524<br />

E.2.2 Tensor<strong>in</strong>nerproduct.....................................524<br />

E.3 Tensorproperties...........................................525<br />

E.4 Contravariantandcovarianttensors ................................526<br />

E.5 Generalized<strong>in</strong>nerproduct......................................527<br />

E.6 Transformationpropertiesofobservables..............................528<br />

F Aspects of multivariate calculus 529<br />

F.1 Partial differentiation ........................................529


xii<br />

CONTENTS<br />

F.2 L<strong>in</strong>earoperators ...........................................529<br />

F.3 TransformationJacobian.......................................531<br />

F.3.1 Transformationof<strong>in</strong>tegrals:.................................531<br />

F.3.2 Transformation of differentialequations:..........................531<br />

F.3.3 PropertiesoftheJacobian: .................................531<br />

F.4 Legendretransformation.......................................532<br />

G Vector differential calculus 533<br />

G.1 Scalar differentialoperators .....................................533<br />

G.1.1 Scalar field ..........................................533<br />

G.1.2 Vector field ..........................................533<br />

G.2 Vector differentialoperators<strong>in</strong>cartesiancoord<strong>in</strong>ates .......................533<br />

G.2.1 Scalar field ..........................................533<br />

G.2.2 Vector field ..........................................534<br />

G.3 Vector differential operators <strong>in</strong> curvil<strong>in</strong>ear coord<strong>in</strong>ates . . ....................535<br />

G.3.1 Gradient: ...........................................535<br />

G.3.2 Divergence: ..........................................536<br />

G.3.3 Curl:..............................................536<br />

G.3.4 Laplacian:...........................................536<br />

H Vector <strong>in</strong>tegral calculus 537<br />

H.1 L<strong>in</strong>e <strong>in</strong>tegral of the gradient of a scalar field............................537<br />

H.2 Divergencetheorem .........................................537<br />

H.2.1 Flux of a vector fieldforGaussiansurface.........................537<br />

H.2.2 Divergence<strong>in</strong>cartesiancoord<strong>in</strong>ates. ............................538<br />

H.3 StokesTheorem............................................540<br />

H.3.1 Thecurl............................................540<br />

H.3.2 Curl<strong>in</strong>cartesiancoord<strong>in</strong>ates................................541<br />

H.4 Potential formulations of curl-free and divergence-free fields ...................543<br />

I Waveform analysis 545<br />

I.1 Harmonicwaveformdecomposition.................................545<br />

I.1.1 PeriodicsystemsandtheFourierseries...........................545<br />

I.1.2 AperiodicsystemsandtheFourierTransform.......................547<br />

I.2 Time-sampledwaveformanalysis ..................................548<br />

I.2.1 Delta-functionimpulseresponse ..............................549<br />

I.2.2 Green’sfunctionwaveformdecomposition .........................550<br />

Bibliography 551<br />

Index 555


Examples<br />

2.1 Example: Explod<strong>in</strong>g cannon shell .................................... 15<br />

2.2 Example: Billiard-ball collisions ..................................... 15<br />

2.3 Example: Bolas thrown by gaucho ................................... 17<br />

2.4 Example: Central force ......................................... 20<br />

2.5 Example: The ideal gas law ....................................... 23<br />

2.6 Example: The mass of galaxies ..................................... 23<br />

2.7 Example: Diatomic molecule ...................................... 26<br />

2.8 Example: Roller coaster ......................................... 27<br />

2.9 Example: Vertical fall <strong>in</strong> the earth’s gravitational field. ........................ 28<br />

2.10 Example: Projectile motion <strong>in</strong> air ................................... 29<br />

2.11 Example: Moment of <strong>in</strong>ertia of a th<strong>in</strong> door .............................. 33<br />

2.12 Example: Merry-go-round ........................................ 33<br />

2.13 Example: Cue pushes a billiard ball .................................. 33<br />

2.14 Example: Center of percussion of a baseball bat ............................ 35<br />

2.15 Example: Energy transfer <strong>in</strong> charged-particle scatter<strong>in</strong>g ....................... 36<br />

2.16 Example: Field of a uniform sphere .................................. 45<br />

3.1 Example: Harmonically-driven series RLC circuit .......................... 65<br />

3.2 Example: Vibration isolation ...................................... 69<br />

3.3 Example: Water waves break<strong>in</strong>g on a beach .............................. 74<br />

3.4 Example: Surface waves for deep water ................................ 74<br />

3.5 Example: Electromagnetic waves <strong>in</strong> ionosphere ............................ 75<br />

3.6 Example: Fourier transform of a Gaussian wave packet: ....................... 77<br />

3.7 Example: Fourier transform of a rectangular wave packet: ...................... 77<br />

3.8 Example: Acoustic wave packet ..................................... 79<br />

3.9 Example: Gravitational red shift .................................... 79<br />

3.10 Example: Quantum baseball ....................................... 80<br />

4.1 Example: Non-l<strong>in</strong>ear oscillator ..................................... 87<br />

5.1 Example: Shortest distance between two po<strong>in</strong>ts ............................110<br />

5.2 Example: Brachistochrone problem ...................................110<br />

5.3 Example: M<strong>in</strong>imal travel cost ......................................112<br />

5.4 Example: Surface area of a cyl<strong>in</strong>drically-symmetric soap bubble ...................113<br />

5.5 Example: Fermat’s Pr<strong>in</strong>ciple ......................................115<br />

5.6 Example: M<strong>in</strong>imum of (∇) 2 <strong>in</strong> a volume ..............................117<br />

5.7 Example: Two dependent variables coupled by one holonomic constra<strong>in</strong>t ..............123<br />

5.8 Example: Catenary ............................................125<br />

5.9 Example: The Queen Dido problem ...................................125<br />

6.1 Example: Motion of a free particle, U=0 ...............................142<br />

6.2 Example: Motion <strong>in</strong> a uniform gravitational field ...........................142<br />

6.3 Example: Central forces .........................................143<br />

6.4 Example: Disk roll<strong>in</strong>g on an <strong>in</strong>cl<strong>in</strong>ed plane ..............................144<br />

6.5 Example: Two connected masses on frictionless <strong>in</strong>cl<strong>in</strong>ed planes ...................147<br />

6.6 Example: Two blocks connected by a frictionless bar .........................148<br />

6.7 Example: Block slid<strong>in</strong>g on a movable frictionless <strong>in</strong>cl<strong>in</strong>ed plane ...................149<br />

6.8 Example: Sphere roll<strong>in</strong>g without slipp<strong>in</strong>g down an <strong>in</strong>cl<strong>in</strong>ed plane on a frictionless floor. .....150<br />

6.9 Example: Mass slid<strong>in</strong>g on a rotat<strong>in</strong>g straight frictionless rod. ....................150<br />

xiii


xiv<br />

EXAMPLES<br />

6.10 Example: Spherical pendulum ......................................151<br />

6.11 Example: Spr<strong>in</strong>g plane pendulum ....................................152<br />

6.12 Example: The yo-yo ...........................................153<br />

6.13 Example: Mass constra<strong>in</strong>ed to move on the <strong>in</strong>side of a frictionless paraboloid ...........154<br />

6.14 Example: Mass on a frictionless plane connected to a plane pendulum ...............155<br />

6.15 Example: Two connected masses constra<strong>in</strong>ed to slide along a mov<strong>in</strong>g rod ..............156<br />

6.16 Example: Mass slid<strong>in</strong>g on a frictionless spherical shell ........................157<br />

6.17 Example: Roll<strong>in</strong>g solid sphere on a spherical shell ..........................159<br />

6.18 Example: Solid sphere roll<strong>in</strong>g plus slipp<strong>in</strong>g on a spherical shell ...................161<br />

6.19 Example: Small body held by friction on the periphery of a roll<strong>in</strong>g wheel ..............162<br />

6.20 Example: Plane pendulum hang<strong>in</strong>g from a vertically-oscillat<strong>in</strong>g support ..............165<br />

6.21 Example: Series-coupled double pendulum subject to impulsive force .................167<br />

7.1 Example: Feynman’s angular-momentum paradox ...........................176<br />

7.2 Example: Atwoods mach<strong>in</strong>e .......................................178<br />

7.3 Example: Conservation of angular momentum for rotational <strong>in</strong>variance: ..............179<br />

7.4 Example: Diatomic molecules and axially-symmetric nuclei .....................180<br />

7.5 Example: L<strong>in</strong>ear harmonic oscillator on a cart mov<strong>in</strong>g at constant velocity ............185<br />

7.6 Example: Isotropic central force <strong>in</strong> a rotat<strong>in</strong>g frame .........................186<br />

7.7 Example: The plane pendulum .....................................187<br />

7.8 Example: Oscillat<strong>in</strong>g cyl<strong>in</strong>der <strong>in</strong> a cyl<strong>in</strong>drical bowl ..........................187<br />

8.1 Example: Motion <strong>in</strong> a uniform gravitational field ...........................201<br />

8.2 Example: One-dimensional harmonic oscillator ............................201<br />

8.3 Example: Plane pendulum ........................................202<br />

8.4 Example: Hooke’s law force constra<strong>in</strong>ed to the surface of a cyl<strong>in</strong>der .................203<br />

8.5 Example: Electron motion <strong>in</strong> a cyl<strong>in</strong>drical magnetron ........................204<br />

8.6 Example: Spherical pendulum us<strong>in</strong>g Hamiltonian mechanics .....................209<br />

8.7 Example: Spherical pendulum us<strong>in</strong>g ( ˙ ˙ ) .....................210<br />

8.8 Example: Spherical pendulum us<strong>in</strong>g ( ˙) ...................211<br />

8.9 Example: S<strong>in</strong>gle particle mov<strong>in</strong>g <strong>in</strong> a vertical plane under the <strong>in</strong>fluence of an <strong>in</strong>verse-square<br />

central force ...............................................212<br />

8.10 Example: Folded cha<strong>in</strong> ..........................................213<br />

8.11 Example: Fall<strong>in</strong>g cha<strong>in</strong> .........................................214<br />

9.1 Example: Gauge <strong>in</strong>variance <strong>in</strong> electromagnetism ...........................229<br />

10.1 Example: Driven, l<strong>in</strong>early-damped, coupled l<strong>in</strong>ear oscillators .....................239<br />

10.2 Example: Kirchhoff’s rules for electrical circuits ...........................240<br />

10.3 Example: The l<strong>in</strong>early-damped, l<strong>in</strong>ear oscillator: ...........................241<br />

11.1 Example: Central force lead<strong>in</strong>g to a circular orbit =2 cos ...................250<br />

11.2 Example: Orbit equation of motion for a free body ..........................252<br />

11.3 Example: L<strong>in</strong>ear two-body restor<strong>in</strong>g force ...............................265<br />

11.4 Example: Inverse square law attractive force ..............................265<br />

11.5 Example: Attractive <strong>in</strong>verse cubic central force ............................266<br />

11.6 Example: Spirall<strong>in</strong>g mass attached by a str<strong>in</strong>g to a hang<strong>in</strong>g mass ..................267<br />

11.7 Example: Two-body scatter<strong>in</strong>g by an <strong>in</strong>verse cubic force .......................273<br />

12.1 Example: Accelerat<strong>in</strong>g spr<strong>in</strong>g plane pendulum .............................291<br />

12.2 Example: Surface of rotat<strong>in</strong>g liquid ...................................293<br />

12.3 Example: The pirouette .........................................294<br />

12.4 Example: Cranked plane pendulum ...................................296<br />

12.5 Example: Nucleon orbits <strong>in</strong> deformed nuclei ..............................297<br />

12.6 Example: Free fall from rest .......................................301<br />

12.7 Example: Projectile fired vertically upwards ..............................301<br />

12.8 Example: Motion parallel to Earth’s surface ..............................301<br />

13.1 Example: Inertia tensor of a solid cube rotat<strong>in</strong>g about the center of mass. .............316<br />

13.2 Example: Inertia tensor of about a corner of a solid cube. ......................317<br />

13.3 Example: Inertia tensor of a hula hoop ................................319<br />

13.4 Example: Inertia tensor of a th<strong>in</strong> book .................................319


EXAMPLES<br />

xv<br />

13.5 Example: Rotation about the center of mass of a solid cube .....................321<br />

13.6 Example: Rotation about the corner of the cube ............................322<br />

13.7 Example: Euler angle transformation ..................................327<br />

13.8 Example: Rotation of a dumbbell ....................................332<br />

13.9 Example: Precession rate for torque-free rotat<strong>in</strong>g symmetric rigid rotor ..............338<br />

13.10Example: Tennis racquet dynamics ...................................341<br />

13.11Example: Rotation of asymmetrically-deformed nuclei ........................342<br />

13.12Example: The Sp<strong>in</strong>n<strong>in</strong>g “Jack” ....................................345<br />

13.13Example: The Tippe Top ........................................346<br />

13.14Example: Tipp<strong>in</strong>g stability of a roll<strong>in</strong>g wheel .............................349<br />

13.15Example: Forces on the bear<strong>in</strong>gs of a rotat<strong>in</strong>g circular disk .....................350<br />

14.1 Example: The Grand Piano .......................................362<br />

14.2 Example: Two coupled l<strong>in</strong>ear oscillators ................................368<br />

14.3 Example: Two equal masses series-coupled by two equal spr<strong>in</strong>gs ...................370<br />

14.4 Example: Two parallel-coupled plane pendula .............................371<br />

14.5 Example: The series-coupled double plane pendula ..........................373<br />

14.6 Example: Three plane pendula; mean-field l<strong>in</strong>ear coupl<strong>in</strong>g ......................374<br />

14.7 Example: Three plane pendula; nearest-neighbor coupl<strong>in</strong>g ......................376<br />

14.8 Example: System of three bodies coupled by six spr<strong>in</strong>gs ........................378<br />

14.9 Example: L<strong>in</strong>ear triatomic molecular CO 2 ......................379<br />

14.10Example: Benzene r<strong>in</strong>g .........................................381<br />

14.11Example: Two l<strong>in</strong>early-damped coupled l<strong>in</strong>ear oscillators .......................388<br />

14.12Example: Collective motion <strong>in</strong> nuclei .................................391<br />

15.1 Example: Check that a transformation is canonical ..........................398<br />

15.2 Example: Angular momentum: .....................................401<br />

15.3 Example: Lorentz force <strong>in</strong> electromagnetism ..............................404<br />

15.4 Example: Wavemotion: .........................................404<br />

15.5 Example: Two-dimensional, anisotropic, l<strong>in</strong>ear oscillator ......................405<br />

15.6 Example: The eccentricity vector ....................................406<br />

15.7 Example: The identity canonical transformation ...........................412<br />

15.8 Example: The po<strong>in</strong>t canonical transformation .............................412<br />

15.9 Example: The exchange canonical transformation ...........................412<br />

15.10Example: Inf<strong>in</strong>itessimal po<strong>in</strong>t canonical transformation .......................412<br />

15.11Example: 1-D harmonic oscillator via a canonical transformation ..................413<br />

15.12Example: Free particle ..........................................417<br />

15.13Example: Po<strong>in</strong>t particle <strong>in</strong> a uniform gravitational field .......................418<br />

15.14Example: One-dimensional harmonic oscillator ...........................419<br />

15.15Example: The central force problem ..................................419<br />

15.16Example: L<strong>in</strong>early-damped, one-dimensional, harmonic oscillator ..................421<br />

15.17Example:Adiabatic<strong>in</strong>varianceforthesimplependulum .......................428<br />

15.18Example: Harmonic oscillator perturbation ..............................430<br />

15.19Example: L<strong>in</strong>dblad resonance <strong>in</strong> planetary and galactic motion ...................431<br />

16.1 Example: Acoustic waves <strong>in</strong> a gas ...................................451<br />

17.1 Example: Muon lifetime .........................................462<br />

17.2 Example: Relativistic Doppler Effect ..................................463<br />

17.3 Example: Tw<strong>in</strong> paradox .........................................463<br />

17.4 Example: Rocket propulsion .......................................466<br />

17.5 Example: Lagrangian for a relativistic free particle ..........................474<br />

17.6 Example: Relativistic particle <strong>in</strong> an external electromagnetic field ..................475<br />

17.7 Example: The Bohr-Sommerfeld hydrogen atom ............................479<br />

A.1 Example: Eigenvalues and eigenvectors of a real symmetric matrix .................504<br />

D.1 Example: Rotation matrix: .......................................517<br />

D.2 Example: Proof that a rotation matrix is orthogonal .........................518<br />

E.1 Example: Displacement gradient tensor ................................524<br />

F.1 Example: Jacobian for transform from cartesian to spherical coord<strong>in</strong>ates ..............531


xvi<br />

EXAMPLES<br />

H.1 Example: Maxwell’s Flux Equations ..................................539<br />

H.2 Example: Buoyancy forces <strong>in</strong> fluids ..................................540<br />

H.3 Example: Maxwell’s circulation equations ...............................542<br />

H.4 Example: Electromagnetic fields: ....................................543<br />

I.1 Example: Fourier transform of a s<strong>in</strong>gle isolated square pulse: ....................548<br />

I.2 Example: Fourier transform of the Dirac delta function: .......................548


Preface<br />

The goal of this book is to <strong>in</strong>troduce the reader to the <strong>in</strong>tellectual beauty, and philosophical implications,<br />

of the fact that nature obeys variational pr<strong>in</strong>ciples plus Hamilton’s Action Pr<strong>in</strong>ciple which underlie the<br />

Lagrangian and Hamiltonian analytical formulations of classical mechanics. These variational methods,<br />

which were developed for classical mechanics dur<strong>in</strong>g the 18 − 19 century, have become the preem<strong>in</strong>ent<br />

formalisms for classical dynamics, as well as for many other branches of modern science and eng<strong>in</strong>eer<strong>in</strong>g.<br />

The ambitious goal of this book is to lead the reader from the <strong>in</strong>tuitive Newtonian vectorial formulation, to<br />

<strong>in</strong>troduction of the more abstract variational pr<strong>in</strong>ciples that underlie Hamilton’s Pr<strong>in</strong>ciple and the related<br />

Lagrangian and Hamiltonian analytical formulations. This culm<strong>in</strong>ates <strong>in</strong> discussion of the contributions of<br />

variational pr<strong>in</strong>ciples to classical mechanics and the development of relativistic and quantum mechanics.<br />

The broad scope of this book attempts to unify the undergraduate physics curriculum by bridg<strong>in</strong>g the<br />

chasm that divides the Newtonian vector-differential formulation, and the <strong>in</strong>tegral variational formulation of<br />

classical mechanics, as well as the correspond<strong>in</strong>g philosophical approaches adopted <strong>in</strong> classical and quantum<br />

mechanics. This book <strong>in</strong>troduces the powerful variational techniques <strong>in</strong> mathematics, and their application to<br />

physics. Application of the concepts of the variational approach to classical mechanics is ideal for illustrat<strong>in</strong>g<br />

the power and beauty of apply<strong>in</strong>g variational pr<strong>in</strong>ciples.<br />

The development of this textbook was <strong>in</strong>fluenced by three textbooks: The <strong>Variational</strong> <strong>Pr<strong>in</strong>ciples</strong> of<br />

<strong>Mechanics</strong> by Cornelius Lanczos (1949) [La49], <strong>Classical</strong> <strong>Mechanics</strong> (1950) by Herbert Goldste<strong>in</strong>[Go50],<br />

and <strong>Classical</strong> Dynamics of Particles and Systems (1965) by Jerry B. Marion[Ma65]. Marion’s excellent<br />

textbook was unusual <strong>in</strong> partially bridg<strong>in</strong>g the chasm between the outstand<strong>in</strong>g graduate texts by Goldste<strong>in</strong><br />

and Lanczos, and a bevy of <strong>in</strong>troductory texts based on Newtonian mechanics that were available at that<br />

time. The present textbook was developed to provide a more modern presentation of the techniques and<br />

philosophical implications of the variational approaches to classical mechanics, with a breadth and depth<br />

close to that provided by Goldste<strong>in</strong> and Lanczos, but <strong>in</strong> a format that better matches the needs of the<br />

undergraduate student. An additional goal is to bridge the gap between classical and modern physics <strong>in</strong> the<br />

undergraduate curriculum. The underly<strong>in</strong>g philosophical approach adopted by this book was espoused by<br />

Galileo Galilei “You cannot teach a man anyth<strong>in</strong>g; you can only help him f<strong>in</strong>d it with<strong>in</strong> himself.”<br />

This book was written <strong>in</strong> support of the physics junior/senior undergraduate course P235W entitled<br />

“<strong>Variational</strong> <strong>Pr<strong>in</strong>ciples</strong> <strong>in</strong> <strong>Classical</strong> <strong>Mechanics</strong>” that the author taught at the University of Rochester between<br />

1993−2015. Initially the lecture notes were distributed to students to allow pre-lecture study, facilitate<br />

accurate transmission of the complicated formulae, and m<strong>in</strong>imize note tak<strong>in</strong>g dur<strong>in</strong>g lectures. These lecture<br />

notes evolved <strong>in</strong>to the present textbook. The target audience of this course typically comprised ≈ 70% junior/senior<br />

undergraduates, ≈ 25% sophomores, ≤ 5% graduate students, and the occasional well-prepared<br />

freshman. The target audience was physics and astrophysics majors, but the course attracted a significant<br />

fraction of majors from other discipl<strong>in</strong>es such as mathematics, chemistry, optics, eng<strong>in</strong>eer<strong>in</strong>g, music, and the<br />

humanities. As a consequence, the book <strong>in</strong>cludes appreciable <strong>in</strong>troductory level physics, plus mathematical<br />

review material, to accommodate the diverse range of prior preparation of the students. This textbook<br />

<strong>in</strong>cludes material that extends beyond what reasonably can be covered dur<strong>in</strong>g a one-term course. This supplemental<br />

material is presented to show the importance and broad applicability of variational concepts to<br />

classical mechanics. The book <strong>in</strong>cludes 161 worked examples, plus 158 assigned problems, to illustrate the<br />

concepts presented. Advanced group-theoretic concepts are m<strong>in</strong>imized to better accommodate the mathematical<br />

skills of the typical undergraduate physics major. To conform with modern literature <strong>in</strong> this field,<br />

this book follows the widely-adopted nomenclature used <strong>in</strong> “<strong>Classical</strong> <strong>Mechanics</strong>” by Goldste<strong>in</strong>[Go50], with<br />

recent additions by Johns[Jo05] and this textbook.<br />

The second edition of this book revised the presentation and <strong>in</strong>cludes recent developments <strong>in</strong> the field.<br />

xvii


xviii<br />

PREFACE<br />

The book is broken <strong>in</strong>to four major sections, the first of which presents a brief historical <strong>in</strong>troduction<br />

(chapter 1), followed by a review of the Newtonian formulation of mechanics plus gravitation (chapter<br />

2), l<strong>in</strong>ear oscillators and wave motion (chapter 3), and an <strong>in</strong>troduction to non-l<strong>in</strong>ear dynamics and chaos<br />

(chapter 4). The second section <strong>in</strong>troduces the variational pr<strong>in</strong>ciples of analytical mechanics that underlie<br />

this book. It <strong>in</strong>cludes an <strong>in</strong>troduction to the calculus of variations (chapter 5), the Lagrangian formulation of<br />

mechanics with applications to holonomic and non-holonomic systems (chapter 6), a discussion of symmetries,<br />

<strong>in</strong>variance, plus Noether’s theorem (chapter 7). This book presents an <strong>in</strong>troduction to the Hamiltonian, the<br />

Hamiltonian formulation of mechanics, the Routhian reduction technique, and a discussion of the subtleties<br />

<strong>in</strong>volved <strong>in</strong> apply<strong>in</strong>g variational pr<strong>in</strong>ciples to variable-mass problems.(Chapter 8). The second edition of<br />

this book presents a unified <strong>in</strong>troduction to Hamiltons Pr<strong>in</strong>ciple, <strong>in</strong>troduces a new approach for apply<strong>in</strong>g<br />

Hamilton’s Pr<strong>in</strong>ciple to systems subject to <strong>in</strong>itial boundary conditions, and discusses how best to exploit the<br />

hierarchy of related formulations based on action, Lagrangian/Hamiltonian, and equations of motion, when<br />

solv<strong>in</strong>g problems subject to symmetries (chapter 9). A consolidated <strong>in</strong>troduction to the application of the<br />

variational approach to nonconservative systems is presented (chapter 10). The third section of the book,<br />

applies Lagrangian and Hamiltonian formulations of classical dynamics to central force problems (chapter 11),<br />

motion <strong>in</strong> non-<strong>in</strong>ertial frames (chapter 12), rigid-body rotation (chapter 13), and coupled l<strong>in</strong>ear oscillators<br />

(chapter 14). The fourth section of the book <strong>in</strong>troduces advanced applications of Hamilton’s Action Pr<strong>in</strong>ciple,<br />

Lagrangian mechanics and Hamiltonian mechanics. These <strong>in</strong>clude Poisson brackets, Liouville’s theorem,<br />

canonical transformations, Hamilton-Jacobi theory, the action-angle technique (chapter 15), and classical<br />

mechanics <strong>in</strong> the cont<strong>in</strong>ua (chapter 16). This is followed by a brief review of the revolution <strong>in</strong> classical<br />

mechanics <strong>in</strong>troduced by E<strong>in</strong>ste<strong>in</strong>’s theory of relativistic mechanics. The extended theory of Lagrangian and<br />

Hamiltonian mechanics is used to apply variational techniques to the Special Theory of Relativity, followed<br />

by a discussion of the use of variational pr<strong>in</strong>ciples <strong>in</strong> the development of the General Theory of Relativity<br />

(chapter 17). The book f<strong>in</strong>ishes with a brief review of the role of variational pr<strong>in</strong>ciples <strong>in</strong> bridg<strong>in</strong>g the gap<br />

between classical mechanics and quantum mechanics, (chapter 18). These advanced topics extend beyond<br />

the typical syllabus for an undergraduate classical mechanics course. They are <strong>in</strong>cluded to stimulate student<br />

<strong>in</strong>terest <strong>in</strong> physics by giv<strong>in</strong>g them a glimpse of the physicsatthesummitthattheyhavealreadystruggled<br />

to climb. This glimpse illustrates the breadth of classical mechanics, and the pivotal role that variational<br />

pr<strong>in</strong>ciples have played <strong>in</strong> the development of classical, relativistic, quantal, and statistical mechanics.<br />

The front cover picture of this book shows a sailplane soar<strong>in</strong>g high above the Italian Alps. This picture<br />

epitomizes the unlimited horizon of opportunities provided when the full dynamic range of variational pr<strong>in</strong>ciples<br />

are applied to classical mechanics. The adjacent pictures of the galaxy, and the skier, represent the wide<br />

dynamic range of applicable topics that span from the orig<strong>in</strong> of the universe, to everyday life. These cover<br />

pictures reflect the beauty and unity of the foundation provided by variational pr<strong>in</strong>ciples to the development<br />

of classical mechanics.<br />

Information regard<strong>in</strong>g the associated P235 undergraduate course at the University of Rochester is available<br />

on the web site at http://www.pas.rochester.edu/~cl<strong>in</strong>e/P235/<strong>in</strong>dex.shtml. Information about the<br />

author is available at the Cl<strong>in</strong>e home web site: http://www.pas.rochester.edu/~cl<strong>in</strong>e/<strong>in</strong>dex.html.<br />

The author thanks Meghan Sarkis who prepared many of the illustrations, Joe Easterly who designed<br />

the book cover plus the webpage, and Moriana Garcia who organized the open-access publication. Andrew<br />

Sifa<strong>in</strong> developed the diagnostic problems <strong>in</strong>cluded <strong>in</strong> the book. The author appreciates the permission,<br />

granted by Professor Struckmeier, to quote his published article on the extended Hamilton-Lagrangian<br />

formalism. The author acknowledges the feedback and suggestions made by many students who have taken<br />

this course, as well as helpful suggestions by his colleagues; Andrew Abrams, Adam Hayes, Connie Jones,<br />

Andrew Melchionna, David Munson, Alice Quillen, Richard Sarkis, James Schneeloch, Steven Torrisi, Dan<br />

Watson, and Frank Wolfs. These lecture notes were typed <strong>in</strong> LATEX us<strong>in</strong>g Scientific WorkPlace (MacKichan<br />

Software, Inc.), while Adobe Illustrator, Photoshop, Orig<strong>in</strong>, Mathematica, and MUPAD, were used to prepare<br />

the illustrations.<br />

Douglas Cl<strong>in</strong>e,<br />

University of Rochester, <strong>2019</strong>


Prologue<br />

Two dramatically different philosophical approaches to science were developed <strong>in</strong> the field of classical mechanics<br />

dur<strong>in</strong>g the 17 - 18 centuries. This time period co<strong>in</strong>cided with the Age of Enlightenment <strong>in</strong> Europe<br />

dur<strong>in</strong>g which remarkable <strong>in</strong>tellectual and philosophical developments occurred. This was a time when both<br />

philosophical and causal arguments were equally acceptable <strong>in</strong> science, <strong>in</strong> contrast with current convention<br />

where there appears to be tacit agreement to discourage use of philosophical arguments <strong>in</strong> science.<br />

Snell’s Law: The genesis of two contrast<strong>in</strong>g philosophical approaches<br />

to science relates back to early studies of the reflection<br />

and refraction of light. The velocity of light <strong>in</strong> a medium of refractive<br />

<strong>in</strong>dex equals = <br />

. Thus a light beam <strong>in</strong>cident at an<br />

angle 1 to the normal of a plane <strong>in</strong>terface between medium 1<br />

and medium 2 isrefractedatanangle 2 <strong>in</strong> medium 2 where the<br />

angles are related by Snell’s Law.<br />

s<strong>in</strong> 1<br />

= 1<br />

= 2<br />

(Snell’s Law)<br />

s<strong>in</strong> 2 2 1<br />

Ibn Sahl of Bagdad (984) first described the refraction of light,<br />

while Snell (1621) derived his law mathematically. Both of these<br />

scientists used the “vectorial approach” where the light velocity <br />

is considered to be a vector po<strong>in</strong>t<strong>in</strong>g <strong>in</strong> the direction of propagation.<br />

Fermat’s Pr<strong>in</strong>ciple: Fermat’s pr<strong>in</strong>ciple of least time (1657),<br />

which is based on the work of Hero of Alexandria (∼ 60) and Ibn<br />

al-Haytham (1021), states that “light travels between two given<br />

po<strong>in</strong>ts along the path of shortest time”. The transit time of a<br />

light beam between two locations and <strong>in</strong> a medium with<br />

position-dependent refractive <strong>in</strong>dex () is given by<br />

=<br />

Z <br />

<br />

= 1 <br />

Z <br />

<br />

()<br />

(Fermat’s Pr<strong>in</strong>ciple)<br />

Fermat’s Pr<strong>in</strong>ciple leads to the derivation of Snell’s Law.<br />

Philosophically the physics underly<strong>in</strong>g the contrast<strong>in</strong>g vectorial<br />

and Fermat’s Pr<strong>in</strong>ciple derivations of Snell’s Law are dramatically<br />

different. The vectorial approach is based on differential relations<br />

between the velocity vectors <strong>in</strong> the two media, whereas Fermat’s<br />

variational approach is based on the fact that the light preferentially<br />

selects a path for which the <strong>in</strong>tegral of the transit time<br />

Figure 1: Vectorial and variational representations<br />

of Snell’s Law for refraction of light.<br />

between the <strong>in</strong>itial location and the f<strong>in</strong>al location is m<strong>in</strong>imized.<br />

That is, the first approach is based on “vectorial mechanics” whereas Fermat’s approach is based on<br />

variational pr<strong>in</strong>ciples <strong>in</strong> that the path between the <strong>in</strong>itial and f<strong>in</strong>al locations is varied to f<strong>in</strong>d the path that<br />

m<strong>in</strong>imizes the transit time. Fermat’s enunciation of variational pr<strong>in</strong>ciples <strong>in</strong> physics played a key role <strong>in</strong> the<br />

historical development, and subsequent exploitation, of the pr<strong>in</strong>ciple of least action <strong>in</strong> analytical formulations<br />

of classical mechanics as discussed below.<br />

xix


xx<br />

PROLOGUE<br />

Newtonian mechanics: Momentum and force are vectors that underlie the Newtonian formulation of<br />

classical mechanics. Newton’s monumental treatise, entitled “Philosophiae Naturalis Pr<strong>in</strong>cipia Mathematica”,<br />

published <strong>in</strong> 1687, established his three universal laws of motion, the universal theory of gravitation,<br />

the derivation of Kepler’s three laws of planetary motion, and the development of calculus. Newton’s three<br />

universal laws of motion provide the most <strong>in</strong>tuitive approach to classical mechanics <strong>in</strong> that they are based on<br />

vector quantities like momentum, and the rate of change of momentum, which are related to force. Newton’s<br />

equation of motion<br />

F = p<br />

(Newton’s equation of motion)<br />

<br />

is a vector differential relation between the <strong>in</strong>stantaneous forces and rate of change of momentum, or equivalent<br />

<strong>in</strong>stantaneous acceleration, all of which are vector quantities. Momentum and force are easy to visualize,<br />

and both cause and effect are embedded <strong>in</strong> Newtonian mechanics. Thus, if all of the forces, <strong>in</strong>clud<strong>in</strong>g the<br />

constra<strong>in</strong>t forces, act<strong>in</strong>g on the system are known, then the motion is solvable for two body systems. The<br />

mathematics for handl<strong>in</strong>g Newton’s “vectorial mechanics” approach to classical mechanics is well established.<br />

Analytical mechanics: <strong>Variational</strong> pr<strong>in</strong>ciples apply to many aspects of our daily life. Typical examples<br />

<strong>in</strong>clude; select<strong>in</strong>g the optimum compromise <strong>in</strong> quality and cost when shopp<strong>in</strong>g, select<strong>in</strong>g the fastest route<br />

to travel from home to work, or select<strong>in</strong>g the optimum compromise to satisfy the disparate desires of the<br />

<strong>in</strong>dividuals compris<strong>in</strong>g a family. <strong>Variational</strong> pr<strong>in</strong>ciples underlie the analytical formulation of mechanics. It<br />

is astonish<strong>in</strong>g that the laws of nature are consistent with variational pr<strong>in</strong>ciples <strong>in</strong>volv<strong>in</strong>g the pr<strong>in</strong>ciple of<br />

least action. M<strong>in</strong>imiz<strong>in</strong>g the action <strong>in</strong>tegral led to the development of the mathematical field of variational<br />

calculus, plus the analytical variational approaches to classical mechanics, by Euler, Lagrange, Hamilton,<br />

and Jacobi.<br />

Leibniz, who was a contemporary of Newton, <strong>in</strong>troduced methods based on a quantity called “vis viva”,<br />

which is Lat<strong>in</strong> for “liv<strong>in</strong>g force” and equals twice the k<strong>in</strong>etic energy. Leibniz believed <strong>in</strong> the philosophy<br />

that God created a perfect world where nature would be thrifty <strong>in</strong> all its manifestations. In 1707, Leibniz<br />

proposed that the optimum path is based on m<strong>in</strong>imiz<strong>in</strong>g the time <strong>in</strong>tegral of the vis viva, which is equivalent<br />

to the action <strong>in</strong>tegral of Lagrangian/Hamiltonian mechanics. In 1744 Euler derived the Leibniz result<br />

us<strong>in</strong>g variational concepts while Maupertuis restated the Leibniz result based on teleological arguments.<br />

The development of Lagrangian mechanics culm<strong>in</strong>ated <strong>in</strong> the 1788 publication of Lagrange’s monumental<br />

treatise entitled “Mécanique Analytique”. Lagrange used d’Alembert’s Pr<strong>in</strong>ciple to derive Lagrangian mechanics<br />

provid<strong>in</strong>g a powerful analytical approach to determ<strong>in</strong>e the magnitude and direction of the optimum<br />

trajectories, plus the associated forces.<br />

The culm<strong>in</strong>ation of the development of analytical mechanics occurred <strong>in</strong> 1834 when Hamilton proposed<br />

his Pr<strong>in</strong>ciple of Least Action, as well as develop<strong>in</strong>g Hamiltonian mechanics which is the premier variational<br />

approach <strong>in</strong> science. Hamilton’s concept of least action is def<strong>in</strong>ed to be the time <strong>in</strong>tegral of the Lagrangian.<br />

Hamilton’s Action Pr<strong>in</strong>ciple (1834) m<strong>in</strong>imizes the action <strong>in</strong>tegral def<strong>in</strong>ed by<br />

=<br />

Z <br />

<br />

(q ˙q)<br />

(Hamilton’s Pr<strong>in</strong>ciple)<br />

In the simplest form, the Lagrangian (q ˙q) equals the difference between the k<strong>in</strong>etic energy and the<br />

potential energy . Hamilton’s Least Action Pr<strong>in</strong>ciple underlies Lagrangian mechanics. This Lagrangian is<br />

a function of generalized coord<strong>in</strong>ates plus their correspond<strong>in</strong>g velocities ˙ . Hamilton also developed<br />

the premier variational approach, called Hamiltonian mechanics, that is based on the Hamiltonian (q p)<br />

which is a function of the fundamental position plus the conjugate momentum variables. In 1843<br />

Jacobi provided the mathematical framework required to fully exploit the power of Hamiltonian mechanics.<br />

Note that the Lagrangian, Hamiltonian, and the action <strong>in</strong>tegral, all are scalar quantities which simplifies<br />

derivation of the equations of motion compared with the vector calculus used by Newtonian mechanics.<br />

Figure 2 presents a philosophical roadmap illustrat<strong>in</strong>g the hierarchy of philosophical approaches based on<br />

Hamilton’s Action Pr<strong>in</strong>ciple, that are available for deriv<strong>in</strong>g the equations of motion of a system. The primary<br />

Stage1 uses Hamilton’s Action functional, = R <br />

<br />

(q ˙q) to derive the Lagrangian, and Hamiltonian<br />

functionals which provide the most fundamental and sophisticated level of understand<strong>in</strong>g. Stage1 <strong>in</strong>volves<br />

specify<strong>in</strong>g all the active degrees of freedom, as well as the <strong>in</strong>teractions <strong>in</strong>volved. Stage2 uses the Lagrangian<br />

or Hamiltonian functionals, derived at Stage1, <strong>in</strong> order to derive the equations of motion for the system of


xxi<br />

Hamilton’s action pr<strong>in</strong>ciple<br />

Stage 1<br />

Stage 2<br />

Hamiltonian<br />

Lagrangian<br />

d’ Alembert’s Pr<strong>in</strong>ciple<br />

Equations of motion<br />

Newtonian mechanics<br />

Stage 3<br />

Solution for motion<br />

Initial conditions<br />

Figure 2: Philosophical road map of the hierarchy of stages <strong>in</strong>volved <strong>in</strong> analytical mechanics. Hamilton’s<br />

Action Pr<strong>in</strong>ciple is the foundation of analytical mechanics. Stage 1 uses Hamilton’s Pr<strong>in</strong>ciple to derive the<br />

Lagranian and Hamiltonian. Stage 2 uses either the Lagrangian or Hamiltonian to derive the equations<br />

of motion for the system. Stage 3 uses these equations of motion to solve for the actual motion us<strong>in</strong>g<br />

the assumed <strong>in</strong>itial conditions. The Lagrangian approach can be derived directly based on d’Alembert’s<br />

Pr<strong>in</strong>ciple. Newtonian mechanics can be derived directly based on Newton’s Laws of Motion. The advantages<br />

and power of Hamilton’s Action Pr<strong>in</strong>ciple are unavailable if the Laws of Motion are derived us<strong>in</strong>g either<br />

d’Alembert’s Pr<strong>in</strong>ciple or Newton’s Laws of Motion.<br />

<strong>in</strong>terest. Stage3 then uses these derived equations of motion to solve for the motion of the system subject to<br />

a given set of <strong>in</strong>itial boundary conditions. Note that Lagrange first derived Lagrangian mechanics based on<br />

d’ Alembert’s Pr<strong>in</strong>ciple, while Newton’s Laws of Motion specify the equations of motion used <strong>in</strong> Newtonian<br />

mechanics.<br />

The analytical approach to classical mechanics appeared contradictory to Newton’s <strong>in</strong>tuitive vectorial<br />

treatment of force and momentum. There is a dramatic difference <strong>in</strong> philosophy between the vectordifferential<br />

equations of motion derived by Newtonian mechanics, which relate the <strong>in</strong>stantaneous force to<br />

the correspond<strong>in</strong>g <strong>in</strong>stantaneous acceleration, and analytical mechanics, where m<strong>in</strong>imiz<strong>in</strong>g the scalar action<br />

<strong>in</strong>tegral <strong>in</strong>volves <strong>in</strong>tegrals over space and time between specified <strong>in</strong>itial and f<strong>in</strong>al states. Analytical mechanics<br />

uses variational pr<strong>in</strong>ciples to determ<strong>in</strong>e the optimum trajectory, from a cont<strong>in</strong>uum of tentative possibilities,<br />

by requir<strong>in</strong>g that the optimum trajectory m<strong>in</strong>imizes the action <strong>in</strong>tegral between specified <strong>in</strong>itial and f<strong>in</strong>al<br />

conditions.<br />

Initially there was considerable prejudice and philosophical opposition to use of the variational pr<strong>in</strong>ciples<br />

approach which is based on the assumption that nature follows the pr<strong>in</strong>ciples of economy. The variational<br />

approach is not <strong>in</strong>tuitive, and thus it was considered to be speculative and “metaphysical”, but it was<br />

tolerated as an efficient tool for exploit<strong>in</strong>g classical mechanics. This opposition to the variational pr<strong>in</strong>ciples<br />

underly<strong>in</strong>g analytical mechanics, delayed full appreciation of the variational approach until the start of the<br />

20 century. As a consequence, the <strong>in</strong>tuitive Newtonian formulation reigned supreme <strong>in</strong> classical mechanics<br />

for over two centuries, even though the remarkable problem-solv<strong>in</strong>g capabilities of analytical mechanics were<br />

recognized and exploited follow<strong>in</strong>g the development of analytical mechanics by Lagrange.<br />

The full significance and superiority of the analytical variational formulations of classical mechanics<br />

became well recognised and accepted follow<strong>in</strong>g the development of the Special Theory of Relativity <strong>in</strong> 1905.<br />

The Theory of Relativity requires that the laws of nature be <strong>in</strong>variant to the reference frame. This is not<br />

satisfied by the Newtonian formulation of mechanics which assumes one absolute frame of reference and a<br />

separation of space and time. In contrast, the Lagrangian and Hamiltonian formulations of the pr<strong>in</strong>ciple of<br />

least action rema<strong>in</strong> valid <strong>in</strong> the Theory of Relativity, if the Lagrangian is written <strong>in</strong> a relativistically-<strong>in</strong>variant


xxii<br />

PROLOGUE<br />

form <strong>in</strong> space-time. The complete <strong>in</strong>variance of the variational approach to coord<strong>in</strong>ate frames is precisely<br />

the formalism necessary for handl<strong>in</strong>g relativistic mechanics.<br />

Hamiltonian mechanics, which is expressed <strong>in</strong> terms of the conjugate variables (q p), relates classical<br />

mechanics directly to the underly<strong>in</strong>g physics of quantum mechanics and quantum field theory. As a consequence,<br />

the philosophical opposition to exploit<strong>in</strong>g variational pr<strong>in</strong>ciples no longer exists, and Hamiltonian<br />

mechanics has become the preem<strong>in</strong>ent formulation of modern physics. The reader is free to draw their own<br />

conclusions regard<strong>in</strong>g the philosophical question “is the pr<strong>in</strong>ciple of economy a fundamental law of classical<br />

mechanics, or is it a fortuitous consequence of the fundamental laws of nature?”<br />

From the late seventeenth century, until the dawn of modern physics at the start of the twentieth century,<br />

classical mechanics rema<strong>in</strong>ed a primary driv<strong>in</strong>g force <strong>in</strong> the development of physics. <strong>Classical</strong> mechanics<br />

embraces an unusually broad range of topics spann<strong>in</strong>g motion of macroscopic astronomical bodies to microscopic<br />

particles <strong>in</strong> nuclear and particle physics, at velocities rang<strong>in</strong>g from zero to near the velocity of<br />

light, from one-body to statistical many-body systems, as well as hav<strong>in</strong>g extensions to quantum mechanics.<br />

Introduction of the Special Theory of Relativity <strong>in</strong> 1905, and the General Theory of Relativity <strong>in</strong> 1916,<br />

necessitated modifications to classical mechanics for relativistic velocities, and can be considered to be an<br />

extended theory of classical mechanics. S<strong>in</strong>ce the 1920 0 s, quantal physics has superseded classical mechanics<br />

<strong>in</strong> the microscopic doma<strong>in</strong>. Although quantum physics has played the lead<strong>in</strong>g role <strong>in</strong> the development of<br />

physics dur<strong>in</strong>g much of the past century, classical mechanics still is a vibrant field of physics that recently<br />

has led to excit<strong>in</strong>g developments associated with non-l<strong>in</strong>ear systems and chaos theory. This has spawned<br />

new branches of physics and mathematics as well as chang<strong>in</strong>g our notion of causality.<br />

Goals: The primary goal of this book is to <strong>in</strong>troduce the reader to the powerful variational-pr<strong>in</strong>ciples<br />

approaches that play such a pivotal role <strong>in</strong> classical mechanics and many other branches of modern science<br />

and eng<strong>in</strong>eer<strong>in</strong>g. This book emphasizes the <strong>in</strong>tellectual beauty of these remarkable developments, as well as<br />

stress<strong>in</strong>g the philosophical implications that have had a tremendous impact on modern science. A secondary<br />

goal is to apply variational pr<strong>in</strong>ciples to solve advanced applications <strong>in</strong> classical mechanics <strong>in</strong> order to<br />

<strong>in</strong>troduce many sophisticated and powerful mathematical techniques that underlie much of modern physics.<br />

This book starts with a review of Newtonian mechanics plus the solutions of the correspond<strong>in</strong>g equations<br />

of motion. This is followed by an <strong>in</strong>troduction to Lagrangian mechanics, based on d’Alembert’s Pr<strong>in</strong>ciple,<br />

<strong>in</strong> order to develop familiarity <strong>in</strong> apply<strong>in</strong>g variational pr<strong>in</strong>ciples to classical mechanics. This leads to <strong>in</strong>troduction<br />

of the more fundamental Hamilton’s Action Pr<strong>in</strong>ciple, plus Hamiltonian mechanics, to illustrate the<br />

power provided by exploit<strong>in</strong>g the full hierarchy of stages available for apply<strong>in</strong>g variational pr<strong>in</strong>ciples to classical<br />

mechanics. F<strong>in</strong>ally the book illustrates how variational pr<strong>in</strong>ciples <strong>in</strong> classical mechanics were exploited<br />

dur<strong>in</strong>g the development of both relativisitic mechanics and quantum physics. The connections and applications<br />

of classical mechanics to modern physics, are emphasized throughout the book <strong>in</strong> an effort to span the<br />

chasm that divides the Newtonian vector-differential formulation, and the <strong>in</strong>tegral variational formulation, of<br />

classical mechanics. This chasm is especially applicable to quantum mechanics which is based completely on<br />

variational pr<strong>in</strong>ciples. Note that variational pr<strong>in</strong>ciples, developed <strong>in</strong> the field of classical mechanics, now are<br />

used <strong>in</strong> a diverse and wide range of fields outside of physics, <strong>in</strong>clud<strong>in</strong>g economics, meteorology, eng<strong>in</strong>eer<strong>in</strong>g,<br />

and comput<strong>in</strong>g.<br />

This study of classical mechanics <strong>in</strong>volves climb<strong>in</strong>g a vast mounta<strong>in</strong> of knowledge, and the pathway to the<br />

top leads to elegant and beautiful theories that underlie much of modern physics. This book exploits variational<br />

pr<strong>in</strong>ciples applied to four major topics <strong>in</strong> classical mechanics to illustrate the power and importance of<br />

variational pr<strong>in</strong>ciples <strong>in</strong> physics. Be<strong>in</strong>g so close to the summit provides the opportunity to take a few extra<br />

steps beyond the normal <strong>in</strong>troductory classical mechanics syllabus to glimpse the excit<strong>in</strong>g physics found at<br />

the summit. This new physics <strong>in</strong>cludes topics such as quantum, relativistic, and statistical mechanics.


Chapter 1<br />

A brief history of classical mechanics<br />

1.1 Introduction<br />

This chapter reviews the historical evolution of classical mechanics s<strong>in</strong>ce considerable <strong>in</strong>sight can be ga<strong>in</strong>ed<br />

from study of the history of science. There are two dramatically different approaches used <strong>in</strong> classical<br />

mechanics. The first is the vectorial approach of Newton which is based on vector quantities like momentum,<br />

force, and acceleration. The second is the analytical approach of Lagrange, Euler, Hamilton, and Jacobi,<br />

that is based on the concept of least action and variational calculus. The more <strong>in</strong>tuitive Newtonian picture<br />

reigned supreme <strong>in</strong> classical mechanics until the start of the twentieth century. <strong>Variational</strong> pr<strong>in</strong>ciples, which<br />

were developed dur<strong>in</strong>g the n<strong>in</strong>eteenth century, never aroused much enthusiasm <strong>in</strong> scientific circles due to<br />

philosophical objections to the underly<strong>in</strong>g concepts;thisapproachwasmerelytoleratedasanefficient tool<br />

for exploit<strong>in</strong>g classical mechanics. A dramatic advance <strong>in</strong> the philosophy of science occurred at the start of<br />

the 20 century lead<strong>in</strong>g to widespread acceptance of the superiority of us<strong>in</strong>g variational pr<strong>in</strong>ciples.<br />

1.2 Greek antiquity<br />

The great philosophers <strong>in</strong> ancient Greece played a key role by us<strong>in</strong>g the astronomical work of the Babylonians<br />

to develop scientific theories of mechanics. Thales of Miletus (624 - 547BC), thefirst of the seven<br />

great greek philosophers, developed geometry, and is hailed as the first true mathematician. Pythagorus<br />

(570 - 495BC) developed mathematics, and postulated that the earth is spherical. Democritus (460 -<br />

370BC) has been called the father of modern science, while Socrates (469 - 399BC) is renowned for his<br />

contributions to ethics. Plato (427-347 B.C.) who was a mathematician and student of Socrates, wrote<br />

important philosophical dialogues. He founded the Academy <strong>in</strong> Athens which was the first <strong>in</strong>stitution of<br />

higher learn<strong>in</strong>g <strong>in</strong> the Western world that helped lay the foundations of Western philosophy and science.<br />

Aristotle (384-322 B.C.) is an important founder of Western philosophy encompass<strong>in</strong>g ethics, logic,<br />

science, and politics. His views on the physical sciences profoundly <strong>in</strong>fluenced medieval scholarship that<br />

extended well <strong>in</strong>to the Renaissance. He presented the first implied formulation of the pr<strong>in</strong>ciple of virtual<br />

work <strong>in</strong> statics, and his statement that “what is lost <strong>in</strong> velocity is ga<strong>in</strong>ed <strong>in</strong> force” is a veiled reference to<br />

k<strong>in</strong>etic and potential energy. He adopted an Earth centered model of the universe. Aristarchus (310 - 240<br />

B.C.) argued that the Earth orbited the Sun and used measurements to imply the relative distances of the<br />

Moon and the Sun. The greek philosophers were relatively advanced <strong>in</strong> logic and mathematics and developed<br />

concepts that enabled them to calculate areas and perimeters. Unfortunately their philosophical approach<br />

neglected collect<strong>in</strong>g quantitative and systematic data that is an essential <strong>in</strong>gredient to the advancement of<br />

science.<br />

Archimedes (287-212 B.C.) represented the culm<strong>in</strong>ation of science <strong>in</strong> ancient Greece. As an eng<strong>in</strong>eer<br />

he designed mach<strong>in</strong>es of war, while as a scientist he made significant contributions to hydrostatics and<br />

the pr<strong>in</strong>ciple of the lever. As a mathematician, he applied <strong>in</strong>f<strong>in</strong>itessimals <strong>in</strong> a way that is rem<strong>in</strong>iscent of<br />

modern <strong>in</strong>tegral calculus, which he used to derive a value for Unfortunately much of the work of the<br />

brilliant Archimedes subsequently fell <strong>in</strong>to oblivion. Hero of Alexandria (10 - 70 A.D.) described the<br />

pr<strong>in</strong>ciple of reflection that light takes the shortest path. This is an early illustration of variational pr<strong>in</strong>ciple<br />

1


2 CHAPTER 1. A BRIEF HISTORY OF CLASSICAL MECHANICS<br />

of least time. Ptolemy (83 - 161 A.D.) wrote several scientific treatises that greatly <strong>in</strong>fluenced subsequent<br />

philosophers. Unfortunately he adopted the <strong>in</strong>correct geocentric solar system <strong>in</strong> contrast to the heliocentric<br />

model of Aristarchus and others.<br />

1.3 Middle Ages<br />

Thedecl<strong>in</strong>eandfalloftheRomanEmpire<strong>in</strong>∼410 A.D. marks the end of <strong>Classical</strong> Antiquity, and the<br />

beg<strong>in</strong>n<strong>in</strong>g of the Dark Ages <strong>in</strong> Western Europe (Christendom), while the Muslim scholars <strong>in</strong> Eastern Europe<br />

cont<strong>in</strong>ued to make progress <strong>in</strong> astronomy and mathematics. For example, <strong>in</strong> Egypt, Alhazen (965 - 1040<br />

A.D.) expanded the pr<strong>in</strong>ciple of least time to reflection and refraction. The Dark Ages <strong>in</strong>volved a long<br />

scientific decl<strong>in</strong>e <strong>in</strong> Western Europe that languished for about 900 years. Science was dom<strong>in</strong>ated by religious<br />

dogma, all western scholars were monks, and the important scientific achievements of Greek antiquity were<br />

forgotten. The works of Aristotle were re<strong>in</strong>troduced to Western Europe by Arabs <strong>in</strong> the early 13 century<br />

lead<strong>in</strong>g to the concepts of forces <strong>in</strong> static systems which were developed dur<strong>in</strong>g the fourteenth century.<br />

This <strong>in</strong>cluded concepts of the work done by a force, and the virtual work <strong>in</strong>volved <strong>in</strong> virtual displacements.<br />

Leonardo da V<strong>in</strong>ci (1452-1519) was a leader <strong>in</strong> mechanics at that time. He made sem<strong>in</strong>al contributions<br />

to science, <strong>in</strong> addition to his well known contributions to architecture, eng<strong>in</strong>eer<strong>in</strong>g, sculpture, and art.<br />

Nicolaus Copernicus (1473-1543) rejected the geocentric theory of Ptolomy and formulated a scientificallybased<br />

heliocentric cosmology that displaced the Earth from the center of the universe. The Ptolomic view<br />

was that heaven represented the perfect unchang<strong>in</strong>g div<strong>in</strong>e while the earth represented change plus chaos,<br />

and the celestial bodies moved relative to the fixed heavens. The book, De revolutionibus orbium coelestium<br />

(On the Revolutions of the Celestial Spheres), published by Copernicus <strong>in</strong> 1543, is regarded as the start<strong>in</strong>g<br />

po<strong>in</strong>t of modern astronomy and the def<strong>in</strong><strong>in</strong>g epiphany that began the Scientific Revolution. ThebookDe<br />

Magnete written <strong>in</strong> 1600 by the English physician William Gilbert (1540-1603) presented the results of<br />

well-planned studies of magnetism and strongly <strong>in</strong>fluenced the <strong>in</strong>tellectual-scientific evolution at that time.<br />

Johannes Kepler (1571-1630), a German mathematician, astronomer and astrologer, was a key<br />

figure <strong>in</strong> the 17th century Scientific Revolution. He is best known for recogniz<strong>in</strong>g the connection between the<br />

motions <strong>in</strong> the sky and physics. His laws of planetary motion were developed by later astronomers based on<br />

his written work Astronomia nova, Harmonices Mundi, andEpitome of Copernican Astrononomy. Kepler<br />

was an assistant to Tycho Brahe (1546-1601) who for many years recorded accurate astronomical data<br />

that played a key role <strong>in</strong> the development of Kepler’s theory of planetary motion. Kepler’s work provided<br />

the foundation for Isaac Newton’s theory of universal gravitation. Unfortunately Kepler did not recognize<br />

the true nature of the gravitational force.<br />

Galileo Galilei (1564-1642) built on the Aristotle pr<strong>in</strong>ciple by recogniz<strong>in</strong>g the law of <strong>in</strong>ertia, the<br />

persistence of motion if no forces act, and the proportionality between force and acceleration. This amounts<br />

to recognition of work as the product of force times displacement <strong>in</strong> the direction of the force. He applied<br />

virtual work to the equilibrium of a body on an <strong>in</strong>cl<strong>in</strong>ed plane. He also showed that the same pr<strong>in</strong>ciple<br />

applies to hydrostatic pressure that had been established by Archimedes, but he did not apply his concepts<br />

<strong>in</strong> classical mechanics to the considerable knowledge base on planetary motion. Galileo is famous for the<br />

apocryphal story that he dropped two cannon balls of differentmassesfromtheTowerofPisatodemonstrate<br />

that their speed of descent was <strong>in</strong>dependent of their mass.<br />

1.4 Age of Enlightenment<br />

The Age of Enlightenment is a term used to describe a phase <strong>in</strong> Western philosophy and cultural life <strong>in</strong><br />

which reason was advocated as the primary source and legitimacy for authority. It developed simultaneously<br />

<strong>in</strong> Germany, France, Brita<strong>in</strong>, the Netherlands, and Italy around the 1650’s and lasted until the French<br />

Revolution <strong>in</strong> 1789. The <strong>in</strong>tellectual and philosophical developments led to moral, social, and political<br />

reforms. The pr<strong>in</strong>ciples of <strong>in</strong>dividual rights, reason, common sense, and deism were a revolutionary departure<br />

from the exist<strong>in</strong>g theocracy, autocracy, oligarchy, aristocracy, and the div<strong>in</strong>e right of k<strong>in</strong>gs. It led to political<br />

revolutions <strong>in</strong> France and the United States. It marks a dramatic departure from the Early Modern period<br />

which was noted for religious authority, absolute state power, guild-based economic systems, and censorship<br />

of ideas. It opened a new era of rational discourse, liberalism, freedom of expression, and scientific method.<br />

This new environment led to tremendous advances <strong>in</strong> both science and mathematics <strong>in</strong> addition to music,


1.4. AGE OF ENLIGHTENMENT 3<br />

literature, philosophy, and art. Scientific development dur<strong>in</strong>g the 17 century <strong>in</strong>cluded the pivotal advances<br />

made by Newton and Leibniz at the beg<strong>in</strong>n<strong>in</strong>g of the revolutionary Age of Enlightenment, culm<strong>in</strong>at<strong>in</strong>g <strong>in</strong> the<br />

development of variational calculus and analytical mechanics by Euler and Lagrange. The scientific advances<br />

of this age <strong>in</strong>clude publication of two monumental books Philosophiae Naturalis Pr<strong>in</strong>cipia Mathematica by<br />

Newton <strong>in</strong> 1687 and Mécanique analytique by Lagrange <strong>in</strong> 1788. These are the def<strong>in</strong>itive two books upon<br />

which classical mechanics is built.<br />

René Descartes (1596-1650) attempted to formulate the laws of motion <strong>in</strong> 1644. He talked about<br />

conservation of motion (momentum) <strong>in</strong> a straight l<strong>in</strong>e but did not recognize the vector character of momentum.<br />

Pierre de Fermat (1601-1665) and René Descartes were two lead<strong>in</strong>g mathematicians <strong>in</strong> the first<br />

half of the 17 century. Independently they discovered the pr<strong>in</strong>ciples of analytic geometry and developed<br />

some <strong>in</strong>itial concepts of calculus. Fermat and Blaise Pascal (1623-1662) were the founders of the theory<br />

of probability.<br />

Isaac Newton (1642-1727) made pioneer<strong>in</strong>g contributions to physics and mathematics as well as<br />

be<strong>in</strong>g a theologian. At 18 he was admitted to Tr<strong>in</strong>ity College Cambridge where he read the writ<strong>in</strong>gs of<br />

modern philosophers like Descartes, and astronomers like Copernicus, Galileo, and Kepler. By 1665 he had<br />

discovered the generalized b<strong>in</strong>omial theorem, and began develop<strong>in</strong>g <strong>in</strong>f<strong>in</strong>itessimal calculus. Due to a plague,<br />

the university closed for two years <strong>in</strong> 1665 dur<strong>in</strong>g which Newton worked at home develop<strong>in</strong>g the theory<br />

of calculus that built upon the earlier work of Barrow and Descartes. He was elected Lucasian Professor<br />

of Mathematics <strong>in</strong> 1669 at the age of 26. From 1670 Newton focussed on optics lead<strong>in</strong>g to his Hypothesis<br />

of Light published <strong>in</strong> 1675 and his book Opticks <strong>in</strong> 1704. Newton described light as be<strong>in</strong>g made up of a<br />

flow of extremely subtle corpuscles that also had associated wavelike properties to expla<strong>in</strong> diffraction and<br />

optical <strong>in</strong>terference that he studied. Newton returned to mechanics <strong>in</strong> 1677 by study<strong>in</strong>g planetary motion<br />

and gravitation that applied the calculus he had developed. In 1687 he published his monumental treatise<br />

entitled Philosophiae Naturalis Pr<strong>in</strong>cipia Mathematica which established his three universal laws of motion,<br />

the universal theory of gravitation, derivation of Kepler’s three laws of planetary motion, and was his first<br />

publication of the development of calculus which he called “the science of fluxions”. Newton’s laws of motion<br />

are based on the concepts of force and momentum, that is, force equals the rate of change of momentum.<br />

Newton’s postulate of an <strong>in</strong>visible force able to act over vast distances led him to be criticized for <strong>in</strong>troduc<strong>in</strong>g<br />

“occult agencies” <strong>in</strong>to science. In a remarkable achievement, Newton completely solved the laws of mechanics.<br />

His theory of classical mechanics and of gravitation reigned supreme until the development of the Theory<br />

of Relativity <strong>in</strong> 1905. The followers of Newton envisioned the Newtonian laws to be absolute and universal.<br />

This dogmatic reverence of Newtonian mechanics prevented physicists from an unprejudiced appreciation of<br />

the analytic variational approach to mechanics developed dur<strong>in</strong>g the 17 through 19 centuries. Newton<br />

was the first scientist to be knighted and was appo<strong>in</strong>ted president of the Royal Society.<br />

Gottfried Leibniz (1646-1716) was a brilliant German philosopher, a contemporary of Newton, who<br />

worked on both calculus and mechanics. Leibniz started development of calculus <strong>in</strong> 1675, ten years after<br />

Newton, but Leibniz published his work <strong>in</strong> 1684, which was three years before Newton’s Pr<strong>in</strong>cipia. Leibniz<br />

made significant contributions to <strong>in</strong>tegral calculus and developed the notation currently used <strong>in</strong> calculus.<br />

He <strong>in</strong>troduced the name calculus based on the Lat<strong>in</strong> word for the small stone used for count<strong>in</strong>g. Newton<br />

and Leibniz were <strong>in</strong>volved <strong>in</strong> a protracted argument over who orig<strong>in</strong>ated calculus. It appears that Leibniz<br />

saw drafts of Newton’s work on calculus dur<strong>in</strong>g a visit to England. Throughout their argument Newton<br />

was the ghost writer of most of the articles <strong>in</strong> support of himself and he had them published under nonde-plume<br />

of his friends. Leibniz made the tactical error of appeal<strong>in</strong>g to the Royal Society to <strong>in</strong>tercede on<br />

his behalf. Newton, as president of the Royal Society, appo<strong>in</strong>ted his friends to an “impartial” committee to<br />

<strong>in</strong>vestigate this issue, then he wrote the committee’s report that accused Leibniz of plagiarism of Newton’s<br />

work on calculus, after which he had it published by the Royal Society. Still unsatisfied he then wrote an<br />

anonymous review of the report <strong>in</strong> the Royal Society’s own periodical. This bitter dispute lasted until the<br />

death of Leibniz. When Leibniz died his work was largely discredited. The fact that he falsely claimed to be<br />

a nobleman and added the prefix “von” to his name, coupled with Newton’s vitriolic attacks, did not help<br />

his credibility. Newton is reported to have declared that he took great satisfaction <strong>in</strong> “break<strong>in</strong>g Leibniz’s<br />

heart.” Studies dur<strong>in</strong>g the 20 century have largely revived the reputation of Leibniz and he is recognized<br />

to have made major contributions to the development of calculus.


4 CHAPTER 1. A BRIEF HISTORY OF CLASSICAL MECHANICS<br />

Figure 1.1: Chronological roadmap of the parallel development of the Newtonian and <strong>Variational</strong>-pr<strong>in</strong>ciples<br />

approaches to classical mechanics.


1.5. VARIATIONAL METHODS IN PHYSICS 5<br />

1.5 <strong>Variational</strong> methods <strong>in</strong> physics<br />

Pierre de Fermat (1601-1665) revived the pr<strong>in</strong>ciple of least time, which states that light travels between<br />

two given po<strong>in</strong>ts along the path of shortest time and was used to derive Snell’s law <strong>in</strong> 1657. This enunciation<br />

of variational pr<strong>in</strong>ciples <strong>in</strong> physics played a key role <strong>in</strong> the historical development of the variational pr<strong>in</strong>ciple<br />

of least action that underlies the analytical formulations of classical mechanics.<br />

Gottfried Leibniz (1646-1716) made significant contributions to the development of variational pr<strong>in</strong>ciples<br />

<strong>in</strong> classical mechanics. In contrast to Newton’s laws of motion, which are based on the concept of<br />

momentum, Leibniz devised a new theory of dynamics based on k<strong>in</strong>etic and potential energy that anticipates<br />

the analytical variational approach of Lagrange and Hamilton. Leibniz argued for a quantity called the “vis<br />

viva”, which is Lat<strong>in</strong> for liv<strong>in</strong>g force, that equals twice the k<strong>in</strong>etic energy. Leibniz argued that the change<br />

<strong>in</strong> k<strong>in</strong>etic energy is equal to the work done. In 1687 Leibniz proposed that the optimum path is based on<br />

m<strong>in</strong>imiz<strong>in</strong>g the time <strong>in</strong>tegral of the vis viva, which is equivalent to the action <strong>in</strong>tegral. Leibniz used both<br />

philosophical and causal arguments <strong>in</strong> his work which were acceptable dur<strong>in</strong>g the Age of Enlightenment. Unfortunately<br />

for Leibniz, his analytical approach based on energies, which are scalars, appeared contradictory<br />

to Newton’s <strong>in</strong>tuitive vectorial treatment of force and momentum. There was considerable prejudice and<br />

philosophical opposition to the variational approach which assumes that nature is thrifty <strong>in</strong> all of its actions.<br />

The variational approach was considered to be speculative and “metaphysical” <strong>in</strong> contrast to the causal<br />

arguments support<strong>in</strong>g Newtonian mechanics. This opposition delayed full appreciation of the variational<br />

approach until the start of the 20 century.<br />

Johann Bernoulli (1667-1748) was a Swiss mathematician who was a student of Leibniz’s calculus, and<br />

sided with Leibniz <strong>in</strong> the Newton-Leibniz dispute over the credit for develop<strong>in</strong>g calculus. Also Bernoulli sided<br />

with the Descartes’ vortex theory of gravitation which delayed acceptance of Newton’s theory of gravitation<br />

<strong>in</strong> Europe. Bernoulli pioneered development of the calculus of variations by solv<strong>in</strong>g the problems of the<br />

catenary, the brachistochrone, and Fermat’s pr<strong>in</strong>ciple. Johann Bernoulli’s son Daniel played a significant<br />

role <strong>in</strong> the development of the well-known Bernoulli Pr<strong>in</strong>ciple <strong>in</strong> hydrodynamics.<br />

Pierre Louis Maupertuis (1698-1759) was a student of Johann Bernoulli and conceived the universal<br />

hypothesis that <strong>in</strong> nature there is a certa<strong>in</strong> quantity called action which is m<strong>in</strong>imized. Although this bold<br />

assumption correctly anticipates the development of the variational approach to classical mechanics, he<br />

obta<strong>in</strong>ed his hypothesis by an entirely <strong>in</strong>correct method. He was a dilettante whose mathematical prowess<br />

was beh<strong>in</strong>d the high standards of that time, and he could not establish satisfactorily the quantity to be<br />

m<strong>in</strong>imized. His teleological 1 argument was <strong>in</strong>fluenced by Fermat’s pr<strong>in</strong>ciple and the corpuscle theory of light<br />

that implied a close connection between optics and mechanics.<br />

Leonhard Euler (1707-1783) was the preem<strong>in</strong>ent Swiss mathematician of the 18 century and was<br />

a student of Johann Bernoulli. Euler developed, with full mathematical rigor, the calculus of variations<br />

follow<strong>in</strong>g <strong>in</strong> the footsteps of Johann Bernoulli. Euler used variational calculus to solve m<strong>in</strong>imum/maximum<br />

isoperimetric problems that had attracted and challenged the early developers of calculus, Newton, Leibniz,<br />

and Bernoulli. Euler also was the first to solve the rigid-body rotation problem us<strong>in</strong>g the three components<br />

of the angular velocity as k<strong>in</strong>ematical variables. Euler became bl<strong>in</strong>d <strong>in</strong> both eyes by 1766 but that did not<br />

h<strong>in</strong>der his prolific output <strong>in</strong> mathematics due to his remarkable memory and mental capabilities. Euler’s<br />

contributions to mathematics are remarkable <strong>in</strong> quality and quantity; for example dur<strong>in</strong>g 1775 he published<br />

one mathematical paper per week <strong>in</strong> spite of be<strong>in</strong>g bl<strong>in</strong>d. Euler implicitly implied the pr<strong>in</strong>ciple of least<br />

action us<strong>in</strong>g vis visa which is not the exact form explicitly developed by Lagrange.<br />

Jean le Rond d’Alembert (1717-1785) was a French mathematician and physicist who had the<br />

clever idea of extend<strong>in</strong>g use of the pr<strong>in</strong>ciple of virtual work from statics to dynamics. d’Alembert’s Pr<strong>in</strong>ciple<br />

rewrites the pr<strong>in</strong>ciple of virtual work <strong>in</strong> the form<br />

X<br />

(F − ṗ )r =0<br />

=1<br />

where the <strong>in</strong>ertial reaction force ṗ is subtracted from the correspond<strong>in</strong>g force F. This extension of the<br />

pr<strong>in</strong>ciple of virtual work applies equally to both statics and dynamics lead<strong>in</strong>g to a s<strong>in</strong>gle variational pr<strong>in</strong>ciple.<br />

Joseph Louis Lagrange (1736-1813) was an Italian mathematician and a student of Leonhard Euler.<br />

In 1788 Lagrange published his monumental treatise on analytical mechanics entitled Mécanique Analytique<br />

1 Teleology is any philosophical account that holds that f<strong>in</strong>al causes exist <strong>in</strong> nature, analogous to purposes found <strong>in</strong> human<br />

actions, nature <strong>in</strong>herently tends toward def<strong>in</strong>ite ends.


6 CHAPTER 1. A BRIEF HISTORY OF CLASSICAL MECHANICS<br />

which <strong>in</strong>troduces his Lagrangian mechanics analytical technique which is based on d’Alembert’s Pr<strong>in</strong>ciple of<br />

Virtual Work. Lagrangian mechanics is a remarkably powerful technique that is equivalent to m<strong>in</strong>imiz<strong>in</strong>g<br />

the action <strong>in</strong>tegral def<strong>in</strong>ed as<br />

=<br />

Z 2<br />

1<br />

The Lagrangian frequently is def<strong>in</strong>ed to be the difference between the k<strong>in</strong>etic energy and potential<br />

energy . His theory only required the analytical form of these scalar quantities. In the preface of his<br />

book he refers modestly to his extraord<strong>in</strong>ary achievements with the statement “The reader will f<strong>in</strong>d no<br />

figures <strong>in</strong> the work. The methods which I set forth do not require either constructions or geometrical or<br />

mechanical reason<strong>in</strong>gs: but only algebraic operations, subject to a regular and uniform rule of procedure.”<br />

Lagrange also <strong>in</strong>troduced the concept of undeterm<strong>in</strong>ed multipliers to handle auxiliary conditions which<br />

plays a vital part of theoretical mechanics. William Hamilton, an outstand<strong>in</strong>g figure <strong>in</strong> the analytical<br />

formulation of classical mechanics, called Lagrange the “Shakespeare of mathematics,” on account of the<br />

extraord<strong>in</strong>ary beauty, elegance, and depth of the Lagrangian methods. Lagrange also pioneered numerous<br />

significant contributions to mathematics. For example, Euler, Lagrange, and d’Alembert developed much of<br />

the mathematics of partial differential equations. Lagrange survived the French Revolution, and, <strong>in</strong> spite of<br />

be<strong>in</strong>g a foreigner, Napoleon named Lagrange to the Legion of Honour and made him a Count of the Empire<br />

<strong>in</strong> 1808. Lagrange was honoured by be<strong>in</strong>g buried <strong>in</strong> the Pantheon.<br />

Carl Friedrich Gauss (1777-1855) was a German child prodigy who made many significant contributions<br />

to mathematics, astronomy and physics. He did not work directly on the variational approach, but<br />

Gauss’s law, the divergence theorem, and the Gaussian statistical distribution are important examples of<br />

concepts that he developed and which feature prom<strong>in</strong>ently <strong>in</strong> classical mechanics as well as other branches<br />

of physics, and mathematics.<br />

Simeon Poisson (1781-1840), was a brilliant mathematician who was a student of Lagrange. He<br />

developed the Poisson statistical distribution as well as the Poisson equation that features prom<strong>in</strong>ently <strong>in</strong><br />

electromagnetic and other field theories. His major contribution to classical mechanics is development, <strong>in</strong><br />

1809, of the Poisson bracket formalism which featured prom<strong>in</strong>ently <strong>in</strong> development of Hamiltonian mechanics<br />

and quantum mechanics.<br />

The zenith <strong>in</strong> development of the variational approach to classical mechanics occurred dur<strong>in</strong>g the 19 <br />

century primarily due to the work of Hamilton and Jacobi.<br />

William Hamilton (1805-1865) was a brilliant Irish physicist, astronomer and mathematician who was<br />

appo<strong>in</strong>ted professor of astronomy at Dubl<strong>in</strong> when he was barely 22 years old. He developed the Hamiltonian<br />

mechanics formalism of classical mechanics which now plays a pivotal role <strong>in</strong> modern classical and quantum<br />

mechanics. He opened an entirely new world beyond the developments of Lagrange. Whereas the Lagrange<br />

equations of motion are complicated second-order differential equations, Hamilton succeeded <strong>in</strong> transform<strong>in</strong>g<br />

them <strong>in</strong>to a set of first-order differential equations with twice as many variables that consider momenta and<br />

their conjugate positions as <strong>in</strong>dependent variables. The differential equations of Hamilton are l<strong>in</strong>ear, have<br />

separated derivatives, and represent the simplest and most desirable form possible for differential equations to<br />

be used <strong>in</strong> a variational approach. Hence the name “canonical variables” given by Jacobi. Hamilton exploited<br />

the d’Alembert pr<strong>in</strong>ciple to give the first exact formulation of the pr<strong>in</strong>ciple of least action which underlies the<br />

variational pr<strong>in</strong>ciples used <strong>in</strong> analytical mechanics. The form derived by Euler and Lagrange employed the<br />

pr<strong>in</strong>ciple <strong>in</strong> a way that applies only for conservative (scleronomic) cases. A significant discovery of Hamilton<br />

is his realization that classical mechanics and geometrical optics can be handled from one unified viewpo<strong>in</strong>t.<br />

In both cases he uses a “characteristic” function that has the property that, by mere differentiation, the<br />

path of the body, or light ray, can be determ<strong>in</strong>ed by the same partial differential equations. This solution is<br />

equivalent to the solution of the equations of motion.<br />

Carl Gustave Jacob Jacobi (1804-1851), a Prussian mathematician and contemporary of Hamilton,<br />

made significant developments <strong>in</strong> Hamiltonian mechanics. He immediately recognized the extraord<strong>in</strong>ary importance<br />

of the Hamiltonian formulation of mechanics. Jacobi developed canonical transformation theory<br />

and showed that the function, used by Hamilton, is only one special case of functions that generate suitable<br />

canonical transformations. He proved that any complete solution of the partial differential equation,<br />

without the specific boundary conditions applied by Hamilton, is sufficient for the complete <strong>in</strong>tegration of<br />

the equations of motion. This greatly extends the usefulness of Hamilton’s partial differential equations.<br />

In 1843 JacobidevelopedboththePoissonbrackets,andtheHamilton-Jacobi, formulations of Hamiltonian<br />

mechanics. The latter gives a s<strong>in</strong>gle, first-order partial differential equation for the action function <strong>in</strong> terms


1.6. THE 20 CENTURY REVOLUTION IN PHYSICS 7<br />

of the generalized coord<strong>in</strong>ates which greatly simplifies solution of the equations of motion. He also derived<br />

a pr<strong>in</strong>ciple of least action for time-<strong>in</strong>dependent cases that had been studied by Euler and Lagrange.<br />

Jacobi developed a superior approach to the variational <strong>in</strong>tegral that, by elim<strong>in</strong>at<strong>in</strong>g time from the <strong>in</strong>tegral,<br />

determ<strong>in</strong>ed the path without say<strong>in</strong>g anyth<strong>in</strong>g about how the motion occurs <strong>in</strong> time.<br />

James Clerk Maxwell (1831-1879) was a Scottish theoretical physicist and mathematician. His most<br />

prom<strong>in</strong>ent achievement was formulat<strong>in</strong>g a classical electromagnetic theory that united previously unrelated<br />

observations, plus equations of electricity, magnetism and optics, <strong>in</strong>to one consistent theory. Maxwell’s<br />

equations demonstrated that electricity, magnetism and light are all manifestations of the same phenomenon,<br />

namely the electromagnetic field. Consequently, all other classic laws and equations of electromagnetism<br />

were simplified cases of Maxwell’s equations. Maxwell’s achievements concern<strong>in</strong>g electromagnetism have<br />

been called the “second great unification <strong>in</strong> physics”. Maxwell demonstrated that electric and magnetic<br />

fields travel through space <strong>in</strong> the form of waves, and at a constant speed of light. In 1864 Maxwell wrote “A<br />

Dynamical Theory of the Electromagnetic Field” which proposed that light was <strong>in</strong> fact undulations <strong>in</strong> the<br />

same medium that is the cause of electric and magnetic phenomena. His work <strong>in</strong> produc<strong>in</strong>g a unified model<br />

of electromagnetism is one of the greatest advances <strong>in</strong> physics. Maxwell, <strong>in</strong> collaboration with Ludwig<br />

Boltzmann (1844-1906), also helped develop the Maxwell—Boltzmann distribution, which is a statistical<br />

means of describ<strong>in</strong>g aspects of the k<strong>in</strong>etic theory of gases. These two discoveries helped usher <strong>in</strong> the era of<br />

modern physics, lay<strong>in</strong>g the foundation for such fields as special relativity and quantum mechanics. Boltzmann<br />

founded the field of statistical mechanics and was an early staunch advocate of the existence of atoms and<br />

molecules.<br />

Henri Po<strong>in</strong>caré (1854-1912) was a French theoretical physicist and mathematician. He was the first to<br />

present the Lorentz transformations <strong>in</strong> their modern symmetric form and discovered the rema<strong>in</strong><strong>in</strong>g relativistic<br />

velocity transformations. Although there is similarity to E<strong>in</strong>ste<strong>in</strong>’s Special Theory of Relativity, Po<strong>in</strong>caré and<br />

Lorentz still believed <strong>in</strong> the concept of the ether and did not fully comprehend the revolutionary philosophical<br />

change implied by E<strong>in</strong>ste<strong>in</strong>. Po<strong>in</strong>caré worked on the solution of the three-body problem <strong>in</strong> planetary motion<br />

and was the first to discover a chaotic determ<strong>in</strong>istic system which laid the foundations of modern chaos<br />

theory. It rejected the long-held determ<strong>in</strong>istic view that if the position and velocities of all the particles are<br />

known at one time, then it is possible to predict the future for all time.<br />

The last two decades of the 19 century saw the culm<strong>in</strong>ation of classical physics and several important<br />

discoveries that led to a revolution <strong>in</strong> science that toppled classical physics from its throne. The end of the<br />

19 century was a time dur<strong>in</strong>g which tremendous technological progress occurred; flight, the automobile,<br />

and turb<strong>in</strong>e-powered ships were developed, Niagara Falls was harnessed for power, etc. Dur<strong>in</strong>g this period,<br />

He<strong>in</strong>rich Hertz (1857-1894) produced electromagnetic waves confirm<strong>in</strong>g their derivation us<strong>in</strong>g Maxwell’s<br />

equations. Simultaneously he discovered the photoelectric effect which was crucial evidence <strong>in</strong> support of<br />

quantum physics. Technical developments, such as photography, the <strong>in</strong>duction spark coil, and the vacuum<br />

pumpplayedasignificant role <strong>in</strong> scientific discoveries made dur<strong>in</strong>g the 1890’s. At the end of the 19 century,<br />

scientists thought that the basic laws were understood and worried that future physics would be <strong>in</strong> the fifth<br />

decimal place; some scientists worried that little was left for them to discover. However, there rema<strong>in</strong>ed a<br />

few, presumed m<strong>in</strong>or, unexpla<strong>in</strong>ed discrepancies plus new discoveries that led to the revolution <strong>in</strong> science<br />

that occurred at the beg<strong>in</strong>n<strong>in</strong>g of the 20 century.<br />

1.6 The 20 century revolution <strong>in</strong> physics<br />

The two greatest achievements of modern physics occurred at the beg<strong>in</strong>n<strong>in</strong>g of the 20 century. The first<br />

was E<strong>in</strong>ste<strong>in</strong>’s development of the Theory of Relativity; the Special Theory of Relativity <strong>in</strong> 1905 and the<br />

General Theory of Relativity <strong>in</strong> 1915. This was followed <strong>in</strong> 1925 by the development of quantum mechanics.<br />

Albert E<strong>in</strong>ste<strong>in</strong> (1879-1955) developed the Special Theory of Relativity <strong>in</strong> 1905 and the General Theory<br />

of Relativity <strong>in</strong> 1915; both of these revolutionary theories had a profound impact on classical mechanics<br />

and the underly<strong>in</strong>g philosophy of physics. The Newtonian formulation of mechanics was shown to be an<br />

approximation that applies only at low velocities, while the General Theory of Relativity superseded Newton’s<br />

Law of Gravitation and expla<strong>in</strong>ed the Equivalence Pr<strong>in</strong>ciple. The Newtonian concepts of an absolute<br />

frame of reference, plus the assumption of the separation of time and space, were shown to be <strong>in</strong>valid at<br />

relativistic velocities. E<strong>in</strong>ste<strong>in</strong>’s postulate that the laws of physics are the same <strong>in</strong> all <strong>in</strong>ertial frames requires<br />

a revolutionary change <strong>in</strong> the philosophy of time, space and reference frames which leads to a breakdown<br />

<strong>in</strong> the Newtonian formalism of classical mechanics. By contrast, the Lagrange and Hamiltonian variational


8 CHAPTER 1. A BRIEF HISTORY OF CLASSICAL MECHANICS<br />

formalisms of mechanics, plus the pr<strong>in</strong>ciple of least action, rema<strong>in</strong> <strong>in</strong>tact us<strong>in</strong>g a relativistically <strong>in</strong>variant<br />

Lagrangian. The <strong>in</strong>dependence of the variational approach to reference frames is precisely the formalism<br />

necessary for relativistic mechanics. The <strong>in</strong>variance to coord<strong>in</strong>ate frames of the basic field equations also<br />

must rema<strong>in</strong> <strong>in</strong>variant for the General Theory of Relativity which also can be derived <strong>in</strong> terms of a relativistic<br />

action pr<strong>in</strong>ciple. Thus the development of the Theory of Relativity unambiguously demonstrated the<br />

superiority of the variational formulation of classical mechanics over the vectorial Newtonian formulation,<br />

and thus the considerable effort made by Euler, Lagrange, Hamilton, Jacobi, and others <strong>in</strong> develop<strong>in</strong>g the<br />

analytical variational formalism of classical mechanics f<strong>in</strong>ally came to fruition at the start of the 20 century.<br />

Newton’s two crown<strong>in</strong>g achievements, the Laws of Motion and the Laws of Gravitation, that had reigned<br />

supreme s<strong>in</strong>ce published <strong>in</strong> the Pr<strong>in</strong>cipia <strong>in</strong> 1687, were toppled from the throne by E<strong>in</strong>ste<strong>in</strong>.<br />

Emmy Noether (1882-1935) has been described as “the greatest ever woman mathematician”. In<br />

1915 she proposed a theorem that a conservation law is associated with any differentiable symmetry of a<br />

physical system. Noether’s theorem evolves naturally from Lagrangian and Hamiltonian mechanics and<br />

she applied it to the four-dimensional world of general relativity. Noether’s theorem has had an important<br />

impact <strong>in</strong> guid<strong>in</strong>g the development of modern physics.<br />

Other profound developments that had revolutionary impacts on classical mechanics were quantum<br />

physics and quantum field theory. The 1913 model of atomic structure by Niels Bohr (1885-1962) and<br />

the subsequent enhancements by Arnold Sommerfeld (1868-1951), were based completely on classical<br />

Hamiltonian mechanics. The proposal of wave-particle duality by Louis de Broglie (1892-1987), made<br />

<strong>in</strong> his 1924 thesis, was the catalyst lead<strong>in</strong>g to the development of quantum mechanics. In 1925 Werner<br />

Heisenberg (1901-1976), and Max Born (1882-1970) developed a matrix representation of quantum<br />

mechanics us<strong>in</strong>g non-commut<strong>in</strong>g conjugate position and momenta variables.<br />

Paul Dirac (1902-1984) showed <strong>in</strong> his Ph.D. thesis that Heisenberg’s matrix representation of quantum<br />

physics is based on the Poisson Bracket generalization of Hamiltonian mechanics, which, <strong>in</strong> contrast to<br />

Hamilton’s canonical equations, allows for non-commut<strong>in</strong>g conjugate variables. In 1926 Erw<strong>in</strong> Schröd<strong>in</strong>ger<br />

(1887-1961) <strong>in</strong>dependently <strong>in</strong>troduced the operational viewpo<strong>in</strong>t and re<strong>in</strong>terpreted the partial differential<br />

equation of Hamilton-Jacobi as a wave equation. His start<strong>in</strong>g po<strong>in</strong>t was the optical-mechanical analogy of<br />

Hamilton that is a built-<strong>in</strong> feature of the Hamilton-Jacobi theory. Schröd<strong>in</strong>ger then showed that the wave<br />

mechanics he developed, and the Heisenberg matrix mechanics, are equivalent representations of quantum<br />

mechanics. In 1928 Dirac developed his relativistic equation of motion for the electron and pioneered the<br />

field of quantum electrodynamics. Dirac also <strong>in</strong>troduced the Lagrangian and the pr<strong>in</strong>ciple of least action to<br />

quantum mechanics, and these ideas were developed <strong>in</strong>to the path-<strong>in</strong>tegral formulation of quantum mechanics<br />

and the theory of electrodynamics by Richard Feynman(1918-1988).<br />

The concepts of wave-particle duality, and quantization of observables, both are beyond the classical<br />

notions of <strong>in</strong>f<strong>in</strong>ite subdivisions <strong>in</strong> classical physics. In spite of the radical departure of quantum mechanics<br />

from earlier classical concepts, the basic feature of the differential equations of quantal physics is their selfadjo<strong>in</strong>t<br />

character which means that they are derivable from a variational pr<strong>in</strong>ciple. Thus both the Theory of<br />

Relativity, and quantum physics are consistent with the variational pr<strong>in</strong>ciple of mechanics, and <strong>in</strong>consistent<br />

with Newtonian mechanics. As a consequence Newtonian mechanics has been dislodged from the throne<br />

it occupied s<strong>in</strong>ce 1687, and the <strong>in</strong>tellectually beautiful and powerful variational pr<strong>in</strong>ciples of analytical<br />

mechanics have been validated.<br />

The 2015 observation of gravitational waves is a remarkable recent confirmation of E<strong>in</strong>ste<strong>in</strong>’s General<br />

Theory of Relativity and the validity of the underly<strong>in</strong>g variational pr<strong>in</strong>ciples <strong>in</strong> physics. Another advance <strong>in</strong><br />

physics is the understand<strong>in</strong>g of the evolution of chaos <strong>in</strong> non-l<strong>in</strong>ear systems that have been made dur<strong>in</strong>g the<br />

past four decades. This advance is due to the availability of computers which has reopened this <strong>in</strong>terest<strong>in</strong>g<br />

branch of classical mechanics, that was pioneered by Henri Po<strong>in</strong>caré about a century ago. Although classical<br />

mechanics is the oldest and most mature branch of physics, there still rema<strong>in</strong> new research opportunities <strong>in</strong><br />

this field of physics.<br />

The focus of this book is to <strong>in</strong>troduce the general pr<strong>in</strong>ciples of the mathematical variational pr<strong>in</strong>ciple<br />

approach, and its applications to classical mechanics. It will be shown that the variational pr<strong>in</strong>ciples, that<br />

were developed <strong>in</strong> classical mechanics, now play a crucial role <strong>in</strong> modern physics and mathematics, plus<br />

many other fields of science and technology.<br />

References:<br />

Excellent sources of <strong>in</strong>formation regard<strong>in</strong>g the history of major players <strong>in</strong> the field of classical mechanics<br />

can be found on Wikipedia, and the book “<strong>Variational</strong> Pr<strong>in</strong>ciple of <strong>Mechanics</strong>” by Lanczos.[La49]


Chapter 2<br />

Review of Newtonian mechanics<br />

2.1 Introduction<br />

It is assumed that the reader has been <strong>in</strong>troduced to Newtonian mechanics applied to one or two po<strong>in</strong>t objects.<br />

This chapter reviews Newtonian mechanics for motion of many-body systems as well as for macroscopic<br />

sized bodies. Newton’s Law of Gravitation also is reviewed. The purpose of this review is to ensure that the<br />

reader has a solid foundation of elementary Newtonian mechanics upon which to build the powerful analytic<br />

Lagrangian and Hamiltonian approaches to classical dynamics.<br />

Newtonian mechanics is based on application of Newton’s Laws of motion which assume that the concepts<br />

of distance, time, and mass, are absolute, that is, motion is <strong>in</strong> an <strong>in</strong>ertial frame. The Newtonian idea of<br />

the complete separation of space and time, and the concept of the absoluteness of time, are violated by the<br />

Theory of Relativity as discussed <strong>in</strong> chapter 17. However, for most practical applications, relativistic effects<br />

are negligible and Newtonian mechanics is an adequate description at low velocities. Therefore chapters<br />

2 − 16 will assume velocities for which Newton’s laws of motion are applicable.<br />

2.2 Newton’s Laws of motion<br />

Newton def<strong>in</strong>ed a vector quantity called l<strong>in</strong>ear momentum p which is the product of mass and velocity.<br />

p = ṙ (2.1)<br />

S<strong>in</strong>ce the mass is a scalar quantity, then the velocity vector ṙ and the l<strong>in</strong>ear momentum vector p are<br />

col<strong>in</strong>ear.<br />

Newton’s laws, expressed <strong>in</strong> terms of l<strong>in</strong>ear momentum, are:<br />

1 Law of <strong>in</strong>ertia: A body rema<strong>in</strong>s at rest or <strong>in</strong> uniform motion unless acted upon by a force.<br />

2 Equation of motion: A body acted upon by a force moves <strong>in</strong> such a manner that the time rate of change<br />

of momentum equals the force.<br />

F = p<br />

(2.2)<br />

<br />

3 Action and reaction: If two bodies exert forces on each other, these forces are equal <strong>in</strong> magnitude and<br />

opposite <strong>in</strong> direction.<br />

Newton’s second law conta<strong>in</strong>s the essential physics relat<strong>in</strong>g the force F and the rate of change of l<strong>in</strong>ear<br />

momentum p.<br />

Newton’s first law, the law of <strong>in</strong>ertia, is a special case of Newton’s second law <strong>in</strong> that if<br />

F = p =0 (2.3)<br />

<br />

then p is a constant of motion.<br />

Newton’s third law also can be <strong>in</strong>terpreted as a statement of the conservation of momentum, that is, for<br />

a two particle system with no external forces act<strong>in</strong>g,<br />

F 12 = −F 21 (2.4)<br />

9


10 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

If the forces act<strong>in</strong>g on two bodies are their mutual action and reaction, then equation 24 simplifies to<br />

F 12 + F 21 = p 1<br />

+ p 2<br />

= (p 1 + p 2 )=0 (2.5)<br />

This implies that the total l<strong>in</strong>ear momentum (P = p 1 + p 2 ) is a constant of motion.<br />

Comb<strong>in</strong><strong>in</strong>g equations 21 and 22 leads to a second-order differential equation<br />

F = p<br />

= r<br />

2 = ¨r (2.6)<br />

2 Note that the force on a body F, and the resultant acceleration a = ¨r are col<strong>in</strong>ear. Appendix 2 gives<br />

explicit expressions for the acceleration a <strong>in</strong> cartesian and curvil<strong>in</strong>ear coord<strong>in</strong>ate systems. The def<strong>in</strong>ition of<br />

forcedependsonthedef<strong>in</strong>ition of the mass . Newton’s laws of motion are obeyed to a high precision for<br />

velocities much less than the velocity of light. For example, recent experiments have shown they are obeyed<br />

with an error <strong>in</strong> the acceleration of ∆ ≤ 5 × 10 −14 2 <br />

2.3 Inertial frames of reference<br />

An <strong>in</strong>ertial frame of reference is one <strong>in</strong> which Newton’s Laws of<br />

motion are valid. It is a non-accelerated frame of reference. An<br />

<strong>in</strong>ertial frame must be homogeneous and isotropic. Physical experiments<br />

can be carried out <strong>in</strong> different <strong>in</strong>ertial reference frames.<br />

The Galilean transformation provides a means of convert<strong>in</strong>g between<br />

two <strong>in</strong>ertial frames of reference mov<strong>in</strong>g at a constant relative<br />

velocity. Consider two reference frames and 0 with 0<br />

mov<strong>in</strong>g with constant velocity V at time Figure 21 shows a<br />

Galilean transformation which can be expressed <strong>in</strong> vector form.<br />

r 0 = r − V (2.7)<br />

0 = <br />

Equation 27 gives the boost, assum<strong>in</strong>g Newton’s hypothesis<br />

that the time is <strong>in</strong>variant to change of <strong>in</strong>ertial frames of reference.<br />

Thetimedifferential of this transformation gives<br />

ṙ 0 = ṙ − V (2.8)<br />

¨r 0 = ¨r<br />

Note that the forces <strong>in</strong> the primed and unprimed <strong>in</strong>ertial frames<br />

are related by<br />

F = p<br />

z<br />

x<br />

O<br />

y<br />

r<br />

Vt<br />

Figure 2.1: Frame 0 mov<strong>in</strong>g with a constant<br />

velocity with respect to frame <br />

at the time .<br />

= ¨r =¨r0 = F 0 (2.9)<br />

Thus Newton’s Laws of motion are <strong>in</strong>variant under a Galilean transformation, that is, the <strong>in</strong>ertial mass is<br />

unchanged under Galilean transformations. If Newton’s laws are valid <strong>in</strong> one <strong>in</strong>ertial frame of reference,<br />

then they are valid <strong>in</strong> any frame of reference <strong>in</strong> uniform motion with respect to the first frame of reference.<br />

This <strong>in</strong>variance is called Galilean <strong>in</strong>variance. There are an <strong>in</strong>f<strong>in</strong>ite number of possible <strong>in</strong>ertial frames all<br />

connected by Galilean transformations.<br />

Galilean <strong>in</strong>variance violates E<strong>in</strong>ste<strong>in</strong>’s Theory of Relativity. In order to satisfy E<strong>in</strong>ste<strong>in</strong>’s postulate<br />

that the laws of physics are the same <strong>in</strong> all <strong>in</strong>ertial frames, as well as satisfy Maxwell’s equations for<br />

electromagnetism, it is necessary to replace the Galilean transformation by the Lorentz transformation. As<br />

will be discussed <strong>in</strong> chapter 17, the Lorentz transformation leads to Lorentz contraction and time dilation both<br />

1−( )<br />

of which are related to the parameter ≡<br />

1<br />

2<br />

where is the velocity of light <strong>in</strong> vacuum. Fortunately,<br />

most situations <strong>in</strong> life <strong>in</strong>volve velocities where ; for example, for a body mov<strong>in</strong>g at 25 000m.p.h.<br />

(11 111 ) which is the escape velocity for a body at the surface of the earth, the factor differs from<br />

unity by about 6810 −10 which is negligible. Relativistic effects are significant only <strong>in</strong> nuclear and particle<br />

physics as well as some exotic conditions <strong>in</strong> astrophysics. Thus, for the purpose of classical mechanics,<br />

usually it is reasonable to assume that the Galilean transformation is valid and is well obeyed under most<br />

practical conditions.<br />

z’<br />

x’<br />

O’<br />

r’<br />

P<br />

y’


2.4. FIRST-ORDER INTEGRALS IN NEWTONIAN MECHANICS 11<br />

2.4 First-order <strong>in</strong>tegrals <strong>in</strong> Newtonian mechanics<br />

A fundamental goal of mechanics is to determ<strong>in</strong>e the equations of motion for an −body system, where<br />

the force F acts on the <strong>in</strong>dividual mass where 1 ≤ ≤ . Newton’s second-order equation of motion,<br />

equation 26 must be solved to calculate the <strong>in</strong>stantaneous spatial locations, velocities, and accelerations for<br />

each mass of an -body system. Both F and ¨r are vectors, each hav<strong>in</strong>g three orthogonal components.<br />

The solution of equation 26 <strong>in</strong>volves <strong>in</strong>tegrat<strong>in</strong>g second-order equations of motion subject to a set of <strong>in</strong>itial<br />

conditions. Although this task appears simple <strong>in</strong> pr<strong>in</strong>ciple, it can be exceed<strong>in</strong>gly complicated for many-body<br />

systems. Fortunately, solution of the motion often can be simplified by exploit<strong>in</strong>g three first-order <strong>in</strong>tegrals<br />

of Newton’s equations of motion, that are related directly to conservation of either the l<strong>in</strong>ear momentum,<br />

angular momentum, or energy of the system. In addition, for the special case of these three first-order<br />

<strong>in</strong>tegrals, the <strong>in</strong>ternal motion of any many-body system can be factored out by a simple transformations <strong>in</strong>to<br />

the center of mass of the system. As a consequence, the follow<strong>in</strong>g three first-order <strong>in</strong>tegrals are exploited<br />

extensively <strong>in</strong> classical mechanics.<br />

2.4.1 L<strong>in</strong>ear Momentum<br />

Newton’s Laws can be written as the differential and <strong>in</strong>tegral forms of the first-order time <strong>in</strong>tegral which<br />

equals the change <strong>in</strong> l<strong>in</strong>ear momentum. That is<br />

F = p <br />

<br />

Z 2<br />

1<br />

F =<br />

Z 2<br />

1<br />

p <br />

=(p 2 − p 1 ) <br />

(2.10)<br />

This allows Newton’s law of motion to be expressed directly <strong>in</strong> terms of the l<strong>in</strong>ear momentum p = ṙ of<br />

each of the 1 bodies <strong>in</strong> the system This first-order time <strong>in</strong>tegral features prom<strong>in</strong>ently <strong>in</strong> classical<br />

mechanics s<strong>in</strong>ce it connects to the important concept of l<strong>in</strong>ear momentum p. Thisfirst-order time <strong>in</strong>tegral<br />

gives that the total l<strong>in</strong>ear momentum is a constant of motion when the sum of the external forces is zero.<br />

2.4.2 Angular momentum<br />

The angular momentum L of a particle with l<strong>in</strong>ear momentum p with respect to an orig<strong>in</strong> from which<br />

the position vector r is measured, is def<strong>in</strong>ed by<br />

L ≡ r × p (2.11)<br />

The torque, or moment of the force N with respect to the same orig<strong>in</strong> is def<strong>in</strong>ed to be<br />

N ≡ r × F (2.12)<br />

where r is the position vector from the orig<strong>in</strong> to the po<strong>in</strong>t where the force F is applied. Note that the<br />

torque N can be written as<br />

N = r × p <br />

(2.13)<br />

<br />

Consider the time differential of the angular momentum, L <br />

<br />

L <br />

<br />

= (r × p )= r <br />

× p + r × p <br />

<br />

(2.14)<br />

However,<br />

r <br />

× p = r <br />

× r <br />

=0 (2.15)<br />

<br />

Equations 213 − 215 canbeusedtowritethefirst-order time <strong>in</strong>tegral for angular momentum <strong>in</strong> either<br />

differential or <strong>in</strong>tegral form as<br />

L <br />

<br />

= r × p <br />

= N <br />

Z 2<br />

1<br />

N =<br />

Z 2<br />

1<br />

L <br />

=(L 2 − L 1 ) <br />

(2.16)<br />

Newton’s Law relates torque and angular momentum about the same axis. When the torque about any axis<br />

is zero then angular momentum about that axis is a constant of motion. If the total torque is zero then the<br />

total angular momentum, as well as the components about three orthogonal axes, all are constants.


12 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

2.4.3 K<strong>in</strong>etic energy<br />

The third first-order <strong>in</strong>tegral, that can be used for solv<strong>in</strong>g the equations of motion, is the first-order spatial<br />

<strong>in</strong>tegral R 2<br />

1 F · r . Note that this spatial <strong>in</strong>tegral is a scalar <strong>in</strong> contrast to the first-order time <strong>in</strong>tegrals for<br />

l<strong>in</strong>ear and angular momenta which are vectors. The work done on a mass by a force F <strong>in</strong> transform<strong>in</strong>g<br />

from condition 1 to 2 is def<strong>in</strong>ed to be<br />

Z 2<br />

[ 12 ] <br />

≡ F · r (2.17)<br />

If F is the net resultant force act<strong>in</strong>g on a particle then the <strong>in</strong>tegrand can be written as<br />

F · r = p <br />

· r v <br />

= <br />

· r <br />

= v <br />

<br />

· v = µ <br />

<br />

1<br />

2 (v · v ) = <br />

2 2 = [ ] <br />

(2.18)<br />

where the k<strong>in</strong>etic energy of a particle is def<strong>in</strong>ed as<br />

1<br />

[ ] <br />

≡ 1 2 2 (2.19)<br />

Thus the work done on the particle , thatis,[ 12 ] <br />

equalsthechange<strong>in</strong>k<strong>in</strong>eticenergyoftheparticleif<br />

there is no change <strong>in</strong> other contributions to the total energy such as potential energy, heat dissipation, etc.<br />

That is<br />

∙ 1<br />

[ 12 ] <br />

=<br />

2 2 2 − 1 2<br />

1¸<br />

2 =[ 2 − 1 ] <br />

(2.20)<br />

Thus the differential, and correspond<strong>in</strong>g first <strong>in</strong>tegral, forms of the k<strong>in</strong>etic energy can be written as<br />

F = Z 2<br />

<br />

F · r =( 2 − 1 ) (2.21)<br />

r 1<br />

If the work done on the particle is positive, then the f<strong>in</strong>al k<strong>in</strong>etic energy 2 1 Especially noteworthy is that<br />

the k<strong>in</strong>etic energy [ ] <br />

is a scalar quantity which makes it simple to use. This first-order spatial <strong>in</strong>tegral is the<br />

foundation of the analytic formulation of mechanics that underlies Lagrangian and Hamiltonian mechanics.<br />

2.5 Conservation laws <strong>in</strong> classical mechanics<br />

Elucidat<strong>in</strong>g the dynamics <strong>in</strong> classical mechanics is greatly simplified when conservation laws are applicable.<br />

In nature, isolated many-body systems frequently conserve one or more of the first-order <strong>in</strong>tegrals for l<strong>in</strong>ear<br />

momentum, angular momentum, and mass/energy. Note that mass and energy are coupled <strong>in</strong> the Theory<br />

of Relativity, but for non-relativistic mechanics the conservation of mass and energy are decoupled. Other<br />

observables such as lepton and baryon numbers are conserved, but these conservation laws usually can be<br />

subsumed under conservation of mass for most problems <strong>in</strong> non-relativistic classical mechanics. The power<br />

of conservation laws <strong>in</strong> calculat<strong>in</strong>g classical dynamics makes it useful to comb<strong>in</strong>e the conservation laws<br />

with the first <strong>in</strong>tegrals for l<strong>in</strong>ear momentum, angular momentum, and work-energy, when solv<strong>in</strong>g problems<br />

<strong>in</strong>volv<strong>in</strong>g Newtonian mechanics. These three conservation laws will be derived assum<strong>in</strong>g Newton’s laws of<br />

motion, however, these conservation laws are fundamental laws of nature that apply well beyond the doma<strong>in</strong><br />

of applicability of Newtonian mechanics.<br />

2.6 Motion of f<strong>in</strong>ite-sized and many-body systems<br />

Elementary presentations <strong>in</strong> classical mechanics discuss motion and forces <strong>in</strong>volv<strong>in</strong>g s<strong>in</strong>gle po<strong>in</strong>t particles.<br />

However, <strong>in</strong> real life, s<strong>in</strong>gle bodies have a f<strong>in</strong>ite size <strong>in</strong>troduc<strong>in</strong>g new degrees of freedom such as rotation and<br />

vibration, and frequently many f<strong>in</strong>ite-sized bodies are <strong>in</strong>volved. A f<strong>in</strong>ite-sized body can be thought of as a<br />

system of <strong>in</strong>teract<strong>in</strong>g particles such as the <strong>in</strong>dividual atoms of the body. The <strong>in</strong>teractions between the parts<br />

of the body can be strong which leads to rigid body motion where the positions of the particles are held<br />

fixed with respect to each other, and the body can translate and rotate. When the <strong>in</strong>teraction between the<br />

bodies is weaker, such as for a diatomic molecule, additional vibrational degrees of relative motion between<br />

the <strong>in</strong>dividual atoms are important. Newton’s third law of motion becomes especially important for such<br />

many-body systems.


2.7. CENTER OF MASS OF A MANY-BODY SYSTEM 13<br />

2.7 Centerofmassofamany-bodysystem<br />

r’ i<br />

A f<strong>in</strong>ite sized body needs a reference po<strong>in</strong>t with respect<br />

to which the motion can be described. For example,<br />

there are 8 corners of a cube that could server as reference<br />

po<strong>in</strong>ts, but the motion of each corner is complicated<br />

if the cube is both translat<strong>in</strong>g and rotat<strong>in</strong>g. The<br />

CM<br />

treatment of the behavior of f<strong>in</strong>ite-sized bodies, or manybody<br />

systems, is greatly simplified us<strong>in</strong>g the concept of<br />

center of mass. Thecenterofmassisaparticularfixed<br />

R<br />

po<strong>in</strong>t <strong>in</strong> the body that has an especially valuable property;<br />

that is, the translational motion of a f<strong>in</strong>ite sized<br />

m<br />

i<br />

body can be treated like that of a po<strong>in</strong>t mass located at<br />

r i<br />

the center of mass. In addition the translational motion<br />

is separable from the rotational-vibrational motion of a<br />

many-body system when the motion is described with<br />

respect to the center of mass. Thus it is convenient at<br />

this juncture to <strong>in</strong>troduce the concept of center of mass<br />

of a many-body system.<br />

For a many-body system, the position vector r ,def<strong>in</strong>ed<br />

relative to the laboratory system, is related to the Figure 2.2: Position vector with respect to the<br />

position vector r 0 with respect to the center of mass, and center of mass.<br />

the center-of-mass location R relative to the laboratory<br />

system. That is, as shown <strong>in</strong> figure 22<br />

r = R + r 0 (2.22)<br />

This vector relation def<strong>in</strong>es the transformation between the laboratory and center of mass systems. For<br />

discrete and cont<strong>in</strong>uous systems respectively, the location of the center of mass is uniquely def<strong>in</strong>ed as be<strong>in</strong>g<br />

where<br />

X<br />

Z<br />

r 0 = r 0 =0<br />

(Center of mass def<strong>in</strong>ition)<br />

Def<strong>in</strong>e the total mass as<br />

<br />

=<br />

X<br />

Z<br />

=<br />

<br />

<br />

<br />

(Total mass)<br />

The average location of the system corresponds to the location of the center of mass s<strong>in</strong>ce 1 P<br />

r 0 =0<br />

that is<br />

1 X<br />

r = R + 1 X<br />

r 0 = R (2.23)<br />

<br />

<br />

<br />

The vector R which describes the location of the center of mass, depends on the orig<strong>in</strong> and coord<strong>in</strong>ate<br />

system chosen. For a cont<strong>in</strong>uous mass distribution the location vector of the center of mass is given by<br />

R = 1 X<br />

r = 1 Z<br />

r (2.24)<br />

<br />

<br />

<br />

The center of mass can be evaluated by calculat<strong>in</strong>g the <strong>in</strong>dividual components along three orthogonal axes.<br />

The center-of-mass frame of reference is def<strong>in</strong>ed as the frame for which the center of mass is stationary.<br />

This frame of reference is especially valuable for elucidat<strong>in</strong>g the underly<strong>in</strong>g physics which <strong>in</strong>volves only the<br />

relative motion of the many bodies. That is, the trivial translational motion of the center of mass frame,<br />

which has no <strong>in</strong>fluence on the relative motion of the bodies, is factored out and can be ignored. For example,<br />

a tennis ball (006) approach<strong>in</strong>g the earth (6 × 10 24 ) with velocity could be treated <strong>in</strong> three frames,<br />

(a) assume the earth is stationary, (b) assume the tennis ball is stationary, or (c) the center-of-mass frame.<br />

The latter frame ignores the center of mass motion which has no <strong>in</strong>fluence on the relative motion of the<br />

tennis ball and the earth. The center of l<strong>in</strong>ear momentum and center of mass coord<strong>in</strong>ate frames are identical<br />

<strong>in</strong> Newtonian mechanics but not <strong>in</strong> relativistic mechanics as described <strong>in</strong> chapter 1743.


14 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

2.8 Total l<strong>in</strong>ear momentum of a many-body system<br />

2.8.1 Center-of-mass decomposition<br />

The total l<strong>in</strong>ear momentum P for a system of particlesisgivenby<br />

P =<br />

X<br />

p = <br />

<br />

X<br />

r (2.25)<br />

<br />

It is convenient to describe a many-body system by a position vector r 0 with respect to the center of mass.<br />

That is,<br />

P =<br />

X<br />

p = <br />

<br />

X<br />

r = R + <br />

<br />

s<strong>in</strong>ce P <br />

r 0 =0as given by the def<strong>in</strong>ition of the center of mass. That is;<br />

r = R + r 0 (2.26)<br />

X<br />

<br />

r 0 = R +0=Ṙ (2.27)<br />

<br />

P = Ṙ (2.28)<br />

Thus the total l<strong>in</strong>ear momentum for a system is the same as the momentum of a s<strong>in</strong>gle particle of mass<br />

= P <br />

located at the center of mass of the system.<br />

2.8.2 Equations of motion<br />

The force act<strong>in</strong>g on particle <strong>in</strong> an -particle many-body system, can be separated <strong>in</strong>to an external force<br />

plus <strong>in</strong>ternal forces f between the particles of the system<br />

F <br />

<br />

F = F +<br />

X<br />

f (2.29)<br />

The orig<strong>in</strong> of the external force is from outside of the system while the <strong>in</strong>ternal force is due to the mutual<br />

<strong>in</strong>teraction between the particles <strong>in</strong> the system. Newton’s Law tells us that<br />

Thustherateofchangeoftotalmomentumis<br />

<br />

6=<br />

ṗ = F = F +<br />

X<br />

f (2.30)<br />

<br />

6=<br />

Ṗ =<br />

X<br />

ṗ =<br />

<br />

X<br />

F +<br />

<br />

X<br />

<br />

X<br />

f (2.31)<br />

<br />

6=<br />

Note that s<strong>in</strong>ce the <strong>in</strong>dices are dummy then<br />

X<br />

<br />

X<br />

f = X <br />

<br />

6=<br />

X<br />

f (2.32)<br />

<br />

6=<br />

Substitut<strong>in</strong>g Newton’s third law f = −f <strong>in</strong>to equation 232 implies that<br />

X<br />

<br />

X<br />

f = X <br />

<br />

6=<br />

X<br />

f = −<br />

<br />

6=<br />

X<br />

<br />

X<br />

f =0 (2.33)<br />

<br />

6=


2.8. TOTAL LINEAR MOMENTUM OF A MANY-BODY SYSTEM 15<br />

which is satisfied only for the case where the summations equal zero. That is, for every <strong>in</strong>ternal force, there<br />

is an equal and opposite reaction force that cancels that <strong>in</strong>ternal force.<br />

Therefore the first-order <strong>in</strong>tegral for l<strong>in</strong>ear momentum can be written <strong>in</strong> differential and <strong>in</strong>tegral forms<br />

as<br />

X<br />

Z 2 X<br />

Ṗ =<br />

F = P 2 − P 1 (2.34)<br />

<br />

F <br />

The reaction of a body to an external force is equivalent to a s<strong>in</strong>gle particle of mass located at the center<br />

of mass assum<strong>in</strong>g that the <strong>in</strong>ternal forces cancel due to Newton’s third law.<br />

Note that the total l<strong>in</strong>ear momentum P is conserved if the net external force F is zero, that is<br />

F = P =0 (2.35)<br />

<br />

Therefore the P of the center of mass is a constant. Moreover, if the component of the force along any<br />

direction be is zero, that is,<br />

F P · be<br />

· be = =0 (2.36)<br />

<br />

then P · be is a constant. This fact is used frequently to solve problems <strong>in</strong>volv<strong>in</strong>g motion <strong>in</strong> a constant force<br />

field. For example, <strong>in</strong> the earth’s gravitational field,themomentumofanobjectmov<strong>in</strong>g<strong>in</strong>vacuum<strong>in</strong>the<br />

vertical direction is time dependent because of the gravitational force, whereas the horizontal component of<br />

momentum is constant if no forces act <strong>in</strong> the horizontal direction.<br />

2.1 Example: Explod<strong>in</strong>g cannon shell<br />

Consider a cannon shell of mass moves along a parabolic trajectory <strong>in</strong> the earths gravitational field.<br />

An <strong>in</strong>ternal explosion, generat<strong>in</strong>g an amount of mechanical energy, blows the shell <strong>in</strong>to two parts. One<br />

part of mass where 1 cont<strong>in</strong>ues mov<strong>in</strong>g along the same trajectory with velocity 0 while the other<br />

part is reduced to rest. F<strong>in</strong>d the velocity of the mass immediately after the explosion.<br />

It is important to remember that the energy release is given <strong>in</strong><br />

the center of mass. If the velocity of the shell immediately before the<br />

explosion is and 0 is the velocity of the part immediately after the<br />

M<br />

explosion, then energy conservation gives that 1 2 2 + = 1 2 02 <br />

The conservation of l<strong>in</strong>ear momentum gives = 0 .Elim<strong>in</strong>at<strong>in</strong>g<br />

v<br />

from these equations gives<br />

s<br />

(1-k)M<br />

0 2<br />

=<br />

[(1 − )]<br />

2.2 Example: Billiard-ball collisions<br />

1<br />

<br />

kM<br />

Explod<strong>in</strong>g cannon shell<br />

A billiard ball with mass and <strong>in</strong>cident velocity collides with an identical stationary ball. Assume that<br />

the balls bounce off each other elastically <strong>in</strong> such a way that the <strong>in</strong>cident ball is deflected at a scatter<strong>in</strong>g angle<br />

to the <strong>in</strong>cident direction. Calculate the f<strong>in</strong>al velocities and of the two balls and the scatter<strong>in</strong>g angle<br />

of the target ball. The conservation of l<strong>in</strong>ear momentum <strong>in</strong> the <strong>in</strong>cident direction , and the perpendicular<br />

direction give<br />

= cos + cos <br />

0= s<strong>in</strong> − s<strong>in</strong> <br />

Energy conservation gives .<br />

<br />

2 2 = 2 2 + 2 <br />

2<br />

Solv<strong>in</strong>gthesethreeequationsgives =90 0 − that is, the balls bounce off perpendicular to each other <strong>in</strong><br />

the laboratory frame. The f<strong>in</strong>al velocities are<br />

v’<br />

= cos <br />

= s<strong>in</strong>


16 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

2.9 Angular momentum of a many-body system<br />

2.9.1 Center-of-mass decomposition<br />

As was the case for l<strong>in</strong>ear momentum, for a many-body system it is possible to separate the angular momentum<br />

<strong>in</strong>to two components. One component is the angular momentum about the center of mass and the<br />

other component is the angular motion of the center of mass about the orig<strong>in</strong> of the coord<strong>in</strong>ate system. This<br />

separation is done by describ<strong>in</strong>g the angular momentum of a many-body system us<strong>in</strong>g a position vector r 0 <br />

with respect to the center of mass plus the vector location R of the center of mass.<br />

The total angular momentum<br />

L =<br />

=<br />

=<br />

X<br />

L =<br />

<br />

r = R + r 0 (2.37)<br />

X<br />

r × p <br />

<br />

X<br />

(R + r 0 ) × <br />

³Ṙ + ṙ<br />

0<br />

´<br />

<br />

X<br />

<br />

i<br />

<br />

hr 0 × ṙ 0 + r 0 × Ṙ + R × ṙ 0 + R × Ṙ<br />

(2.38)<br />

Note that if the position vectors are with respect to the center of mass, then P <br />

r 0 =0result<strong>in</strong>g <strong>in</strong> the<br />

middle two terms <strong>in</strong> the bracket be<strong>in</strong>g zero, that is;<br />

L =<br />

X<br />

r 0 × p 0 + R × P (2.39)<br />

<br />

The total angular momentum separates <strong>in</strong>to two terms, the angular momentum about the center of mass,<br />

plus the angular momentum of the center of mass about the orig<strong>in</strong> of the axis system. This factor<strong>in</strong>g of the<br />

angular momentum only applies for the center of mass. This is called Samuel König’s first theorem.<br />

2.9.2 Equations of motion<br />

The time derivative of the angular momentum<br />

˙L = r × p = ṙ × p + r × ṗ (2.40)<br />

But<br />

ṙ × p = ṙ × ṙ =0 (2.41)<br />

Thus the torque act<strong>in</strong>g on mass is given by<br />

N = ˙L = r × ṗ = r × F (2.42)<br />

Consider that the resultant force act<strong>in</strong>g on particle <strong>in</strong> this -particle system can be separated <strong>in</strong>to an<br />

external force F <br />

plus <strong>in</strong>ternal forces between the particles of the system<br />

X<br />

F = F + f (2.43)<br />

The orig<strong>in</strong> of the external force is from outside of the system while the <strong>in</strong>ternal force is due to the <strong>in</strong>teraction<br />

with the other − 1 particles <strong>in</strong> the system. Newton’s Law tells us that<br />

<br />

6=<br />

ṗ = F = F +<br />

X<br />

f (2.44)<br />

<br />

6=


2.9. ANGULAR MOMENTUM OF A MANY-BODY SYSTEM 17<br />

The rate of change of total angular momentum is<br />

˙L = X ˙L = X r × ṗ = X<br />

<br />

<br />

<br />

r × F <br />

+ X <br />

X<br />

r × f (2.45)<br />

<br />

6=<br />

S<strong>in</strong>ce f = −f the last expression can be written as<br />

X X<br />

r × f = X<br />

<br />

<br />

<br />

6=<br />

X<br />

(r − r ) × f (2.46)<br />

<br />

<br />

Note that (r − r ) is the vector r connect<strong>in</strong>g to . For central forces the force vector f = cr thus<br />

X X<br />

(r − r ) × f = X X<br />

r × cr =0 (2.47)<br />

<br />

<br />

<br />

<br />

That is, for central <strong>in</strong>ternal forces the total <strong>in</strong>ternal torque on a system of particles is zero, and the rate of<br />

change of total angular momentum for central <strong>in</strong>ternal forces becomes<br />

<br />

<br />

˙L = X <br />

r × F <br />

= X <br />

N = N (2.48)<br />

where N is the net external torque act<strong>in</strong>g on the system. Equation 248 leads to the differential and <strong>in</strong>tegral<br />

forms of the first <strong>in</strong>tegral relat<strong>in</strong>g the total angular momentum to total external torque.<br />

Z2<br />

˙L = N <br />

1<br />

N = L 2 − L 1 (2.49)<br />

Angular momentum conservation occurs <strong>in</strong> many problems <strong>in</strong>volv<strong>in</strong>g zero external torques N =0 plus<br />

two-body central forces F =()ˆr s<strong>in</strong>ce the torque on the particle about the center of the force is zero<br />

N = r × F =()[r × ˆr] =0 (2.50)<br />

Examples are, the central gravitational force for stellar or planetary systems <strong>in</strong> astrophysics, and the central<br />

electrostatic force manifest for motion of electrons <strong>in</strong> the atom. In addition, the component of angular<br />

momentum about any axis Lê is conserved if the net external torque about that axis Nê =0.<br />

2.3 Example: Bolas thrown by gaucho<br />

Consider the bolas thrown by a gaucho to catch cattle. This is a<br />

system with conserved l<strong>in</strong>ear and angular momentum about certa<strong>in</strong><br />

axes. When the bolas leaves the gaucho’s hand the center of mass<br />

has a l<strong>in</strong>ear velocity V plus an angular momentum about the center<br />

of mass of L If no external torques act, then the center of mass of<br />

the bolas will follow a typical ballistic trajectory <strong>in</strong> the earth’s gravitational<br />

field while the angular momentum vector L is conserved,<br />

that is, both <strong>in</strong> magnitude and direction. The tension <strong>in</strong> the ropes<br />

connect<strong>in</strong>g the three balls does not impact the motion of the system<br />

as long as the ropes do not snap due to centrifugal forces.<br />

CM<br />

V 0<br />

m 1<br />

Bolas thrown by a gaucho


18 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

2.10 Work and k<strong>in</strong>etic energy for a many-body system<br />

2.10.1 Center-of-mass k<strong>in</strong>etic energy<br />

For a many-body system the position vector r 0 with respect to the center of mass is given by.<br />

r = R + r 0 (2.51)<br />

The location of the center of mass is uniquely def<strong>in</strong>ed as be<strong>in</strong>g at the location where R r 0 =0 The<br />

velocity of the particle can be expressed <strong>in</strong> terms of the velocity of the center of mass Ṙ plus the velocity<br />

of the particle with respect to the center of mass ṙ 0 .Thatis,<br />

The total k<strong>in</strong>etic energy is<br />

=<br />

X<br />

<br />

1<br />

2 2 =<br />

X<br />

<br />

1<br />

2 ṙ · ṙ =<br />

X<br />

<br />

ṙ = Ṙ + ṙ 0 (2.52)<br />

Ã<br />

1<br />

2 ṙ 0 · ṙ 0 <br />

+<br />

<br />

!<br />

X<br />

ṙ 0 · Ṙ + X<br />

<br />

<br />

1<br />

2 Ṙ · Ṙ (2.53)<br />

P<br />

For the special case of the center of mass, the middle term is zero s<strong>in</strong>ce, by def<strong>in</strong>ition of the center of mass,<br />

ṙ 0 =0 Therefore X 1<br />

=<br />

2 02 + 1 2 2 (2.54)<br />

<br />

Thus the total k<strong>in</strong>etic energy of the system is equal to the sum of the k<strong>in</strong>etic energy of a mass mov<strong>in</strong>g<br />

with the center of mass velocity plus the k<strong>in</strong>etic energy of motion of the <strong>in</strong>dividual particles relative to the<br />

center of mass. This is called Samuel König’s second theorem.<br />

Note that for a fixed center-of-mass energy, the total k<strong>in</strong>etic energy has a m<strong>in</strong>imum value of P <br />

1 2 <br />

02<br />

when the velocity of the center of mass =0. For a given <strong>in</strong>ternal excitation energy, the m<strong>in</strong>imum energy<br />

required to accelerate collid<strong>in</strong>g bodies occurs when the collid<strong>in</strong>g bodies have identical, but opposite, l<strong>in</strong>ear<br />

momenta. That is, when the center-of-mass velocity =0.<br />

2.10.2 Conservative forces and potential energy<br />

In general, the l<strong>in</strong>e <strong>in</strong>tegral of a force field F, thatis, R 2<br />

F·r is both path and time dependent. However,<br />

1<br />

an important class of forces, called conservative forces, exist for which the follow<strong>in</strong>g two facts are obeyed.<br />

1) Time <strong>in</strong>dependence:<br />

The force depends only on the particle position r, that is, it does not depend on velocity or time.<br />

2) Path <strong>in</strong>dependence:<br />

For any two po<strong>in</strong>ts 1 and 2 , the work done by F is <strong>in</strong>dependent of the path taken between 1 and 2 .<br />

If forces are path <strong>in</strong>dependent, then it is possible to def<strong>in</strong>e a scalar field, called potential energy, denoted<br />

by (r) that is only a function of position. The path <strong>in</strong>dependence can be expressed by not<strong>in</strong>g that the<br />

<strong>in</strong>tegral around a closed loop is zero. That is<br />

I<br />

F · r =0 (2.55)<br />

Apply<strong>in</strong>g Stokes theorem for a path-<strong>in</strong>dependent force leads to the alternate statement that the curl is zero.<br />

See appendix 33<br />

∇ × F =0 (2.56)<br />

Note that the vector product of two del operators ∇ act<strong>in</strong>g on a scalar field equals<br />

∇ × ∇ =0 (2.57)<br />

Thus it is possible to express a path-<strong>in</strong>dependent force field as the gradient of a scalar field, , thatis<br />

F = −∇ (2.58)


2.10. WORK AND KINETIC ENERGY FOR A MANY-BODY SYSTEM 19<br />

Then the spatial <strong>in</strong>tegral<br />

Z 2<br />

1<br />

F · r = −<br />

Z 2<br />

1<br />

(∇) · r = 1 − 2 (2.59)<br />

Thus for a path-<strong>in</strong>dependent force, the work done on the particle is given by the change <strong>in</strong> potential energy<br />

if there is no change <strong>in</strong> k<strong>in</strong>etic energy. For example, if an object is lifted aga<strong>in</strong>st the gravitational field, then<br />

work is done on the particle and the f<strong>in</strong>al potential energy 2 exceeds the <strong>in</strong>itial potential energy, 1 .<br />

2.10.3 Total mechanical energy<br />

The total mechanical energy of a particle is def<strong>in</strong>ed as the sum of the k<strong>in</strong>etic and potential energies.<br />

= + (2.60)<br />

Note that the potential energy is def<strong>in</strong>ed only to with<strong>in</strong> an additive constant s<strong>in</strong>ce the force F = −∇<br />

depends only on difference <strong>in</strong> potential energy. Similarly, the k<strong>in</strong>etic energy is not absolute s<strong>in</strong>ce any <strong>in</strong>ertial<br />

frame of reference can be used to describe the motion and the velocity of a particle depends on the relative<br />

velocities of <strong>in</strong>ertial frames. Thus the total mechanical energy = + is not absolute.<br />

If a s<strong>in</strong>gle particle is subject to several path-<strong>in</strong>dependent forces, such as gravity, l<strong>in</strong>ear restor<strong>in</strong>g forces,<br />

etc., then a potential energy canbeascribedtoeachofthe forces where for each force F = −∇ .In<br />

X<br />

contrast to the forces, which add vectorially, these scalar potential energies are additive, = .Thus<br />

the total mechanical energy for potential energies equals<br />

<br />

= + (r) = +<br />

X<br />

(r) (2.61)<br />

<br />

Thetimederivativeofthetotalmechanicalenergy = + equals<br />

<br />

= <br />

+ <br />

<br />

Equation 218 gave that = F · r. Thus,thefirst term <strong>in</strong> equation 262 equals<br />

(2.62)<br />

<br />

= F · r<br />

(2.63)<br />

<br />

The potential energy can be a function of both position and time. Thus the time difference <strong>in</strong> potential<br />

energy due to change <strong>in</strong> both time and position is given as<br />

<br />

= X <br />

<br />

+ <br />

<br />

r<br />

=(∇) ·<br />

+ <br />

<br />

(2.64)<br />

The time derivative of the total mechanical energy is given us<strong>in</strong>g equations 263 264 <strong>in</strong> equation 262<br />

<br />

= <br />

+ <br />

= F · r r<br />

+(∇) ·<br />

+ <br />

<br />

r<br />

=[F +(∇)] ·<br />

+ <br />

<br />

Note that if the field is path <strong>in</strong>dependent, that is ∇ × F =0 then the force and potential are related by<br />

(2.65)<br />

F = −∇ (2.66)<br />

Therefore, for path <strong>in</strong>dependent forces, the first term <strong>in</strong> the time derivative of the total energy <strong>in</strong> equation<br />

265 is zero. That is,<br />

<br />

= <br />

(2.67)<br />

<br />

In addition, when the potential energy is not an explicit function of time, then <br />

<br />

=0and thus the total<br />

energy is conserved. That is, for the comb<strong>in</strong>ation of (a) path <strong>in</strong>dependence plus (b) time <strong>in</strong>dependence, then<br />

the total energy of a conservative field is conserved.


20 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

Note that there are cases where the concept of potential still is useful even when it is time dependent.<br />

That is, if path <strong>in</strong>dependence applies, i.e. F = −∇ at any <strong>in</strong>stant. For example, a Coulomb field problem<br />

where charges are slowly chang<strong>in</strong>g due to leakage etc., or dur<strong>in</strong>g a peripheral collision between two charged<br />

bodies such as nuclei.<br />

2.4 Example: Central force<br />

A particle of mass moves along a trajectory given by = 0 cos 1 and = 0 s<strong>in</strong> 2 .<br />

a) F<strong>in</strong>d the and components of the force and determ<strong>in</strong>e the condition for which the force is a central<br />

force.<br />

Differentiat<strong>in</strong>g with respect to time gives<br />

˙ = − 0 1 s<strong>in</strong> ( 1 ) ¨ = − 0 2 1 cos ( 1 )<br />

˙ = − 0 2 cos ( 2 ) ¨ = − 0 2 2 s<strong>in</strong> ( 2 )<br />

Newton’s second law gives<br />

F= (¨ˆ+¨ˆ) =− £ 0 2 1 cos ( 1 )ˆ + 0 2 2 s<strong>in</strong> ( 2 ) ˆ ¤ = − £ 2 1ˆ + 2 2ˆ ¤<br />

Note that if 1 = 2 = then<br />

F ==− 2 [ˆ + ˆ] =− 2 r<br />

That is, it is a central force if 1 = 2 = <br />

b) F<strong>in</strong>d the potential energy as a function of and .<br />

S<strong>in</strong>ce<br />

∙ ˆ¸<br />

F = −∇ = − +<br />

ˆ<br />

<br />

then<br />

= 1 2 ¡ 2 1 2 + 2 2 2¢<br />

assum<strong>in</strong>g that =0at the orig<strong>in</strong>.<br />

c) Determ<strong>in</strong>e the k<strong>in</strong>etic energy of the particle and show that it is conserved.<br />

The total energy<br />

= + = 1 2 ¡ ˙ 2 ¢ 1<br />

+ ˙ 2 +<br />

2 ¡ 2 1 2 + 2 2 2¢ = 1 2 ¡ 2 0 2 1 + 0 2 2 ¢<br />

2<br />

s<strong>in</strong>ce cos 2 +s<strong>in</strong> 2 =1. Thus the total energy is a constant and is conserved.<br />

2.10.4 Total mechanical energy for conservative systems<br />

Equation 220 showed that, us<strong>in</strong>g Newton’s second law, F = p<br />

<br />

the first-order spatial <strong>in</strong>tegral gives that<br />

the work done, 12 is related to the change <strong>in</strong> the k<strong>in</strong>etic energy. That is,<br />

12 ≡<br />

Z 2<br />

1<br />

F · r = 1 2 2 2 − 1 2 2 1 = 2 − 1 (2.68)<br />

The work done 12 also can be evaluated <strong>in</strong> terms of the known forces F <strong>in</strong> the spatial <strong>in</strong>tegral.<br />

Consider that the resultant force act<strong>in</strong>g on particle <strong>in</strong> this -particle system can be separated <strong>in</strong>to an<br />

external force F <br />

plus <strong>in</strong>ternal forces between the particles of the system<br />

X<br />

F = F + f (2.69)<br />

The orig<strong>in</strong> of the external force is from outside of the system while the <strong>in</strong>ternal force is due to the <strong>in</strong>teraction<br />

with the other − 1 particles <strong>in</strong> the system. Newton’s Law tells us that<br />

X<br />

ṗ = F = F + f (2.70)<br />

<br />

6=<br />

<br />

6=


2.10. WORK AND KINETIC ENERGY FOR A MANY-BODY SYSTEM 21<br />

The work done on the system by a force mov<strong>in</strong>g from configuration 1 → 2 is given by<br />

1→2 =<br />

X<br />

Z 2<br />

1<br />

F · r +<br />

X<br />

<br />

X<br />

<br />

6=<br />

Z 2<br />

1<br />

f · r (2.71)<br />

S<strong>in</strong>ce f = −f then<br />

1→2 =<br />

X<br />

<br />

Z 2<br />

1<br />

F · r +<br />

X<br />

<br />

X<br />

<br />

<br />

Z 2<br />

1<br />

f · (r − r ) (2.72)<br />

where r − r = r is the vector from to <br />

Assume that both the external and <strong>in</strong>ternal forces are conservative, and thus can be derived from time<br />

<strong>in</strong>dependent potentials, that is<br />

F = −∇ <br />

(2.73)<br />

f = −∇ <br />

(2.74)<br />

Then<br />

1→2 = −<br />

=<br />

X<br />

<br />

X<br />

<br />

Z 2<br />

1<br />

∇ <br />

<br />

<br />

(1) −<br />

Def<strong>in</strong>e the total external potential energy,<br />

and the total <strong>in</strong>ternal energy<br />

X<br />

<br />

· r −<br />

X<br />

<br />

<br />

(2) +<br />

X<br />

<br />

<br />

X<br />

<br />

Z 2<br />

1<br />

∇ <br />

<br />

<br />

(1) −<br />

· r <br />

X<br />

<br />

<br />

(2)<br />

= (1) − (2) + (1) − (2) (2.75)<br />

=<br />

=<br />

Equat<strong>in</strong>g the two equivalent equations for 1→2 ,thatis268 and 275gives that<br />

Regroup these terms <strong>in</strong> equation 278 gives<br />

X<br />

<br />

X<br />

<br />

<br />

(2.76)<br />

<br />

(2.77)<br />

1→2 = 2 − 1 = (1) − (2) + (1) − (2) (2.78)<br />

1 + (1) + (1) = 2 + (2) + (2)<br />

This shows that, for conservative forces, the total energy is conserved and is given by<br />

= + + (2.79)<br />

The three first-order <strong>in</strong>tegrals for l<strong>in</strong>ear momentum, angular momentum, and energy provide powerful<br />

approaches for solv<strong>in</strong>g the motion of Newtonian systems due to the applicability of conservation laws for the<br />

correspond<strong>in</strong>g l<strong>in</strong>ear and angular momentum plus energy conservation for conservative forces. In addition,<br />

the important concept of center-of-mass motion naturally separates out for these three first-order <strong>in</strong>tegrals.<br />

Although these conservation laws were derived assum<strong>in</strong>g Newton’s Laws of motion, these conservation laws<br />

are more generally applicable, and these conservation laws surpass the range of validity of Newton’s Laws of<br />

motion. For example, <strong>in</strong> 1930 Pauli and Fermi postulated the existence of the neutr<strong>in</strong>o <strong>in</strong> order to account for<br />

non-conservation of energy and momentum <strong>in</strong> -decay because they did not wish to rel<strong>in</strong>quish the concepts<br />

of energy and momentum conservation. The neutr<strong>in</strong>o was first detected <strong>in</strong> 1956 confirm<strong>in</strong>g the correctness<br />

of this hypothesis.


22 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

2.11 Virial Theorem<br />

The Virial theorem is an important theorem for a system of mov<strong>in</strong>g particles both <strong>in</strong> classical physics and<br />

quantum physics. The Virial Theorem is useful when consider<strong>in</strong>g a collection of many particles and has a<br />

special importance to central-force motion. For a general system of mass po<strong>in</strong>ts with position vectors r and<br />

applied forces F , consider the scalar product <br />

≡ X <br />

p · r (2.80)<br />

where sums over all particles. The time derivative of is<br />

<br />

= X <br />

p · ṙ + X <br />

ṗ · r (2.81)<br />

However,<br />

X<br />

p · ṙ = X<br />

<br />

<br />

ṙ · ṙ = X <br />

2 =2 (2.82)<br />

Also, s<strong>in</strong>ce ṗ = F <br />

X<br />

<br />

ṗ · r = X <br />

F · r (2.83)<br />

Thus<br />

<br />

=2 + X <br />

F · r (2.84)<br />

Thetimeaverageoveraperiod is<br />

Z<br />

1 <br />

0<br />

() − (0)<br />

= = h2 i +<br />

<br />

* X<br />

<br />

F · r <br />

+<br />

(2.85)<br />

where the hi brackets refer to the time average. Note that if the motion is periodic and the chosen time <br />

equals a multiple of the period, then ()−(0)<br />

<br />

=0. Even if the motion is not periodic, if the constra<strong>in</strong>ts and<br />

velocities of all the particles rema<strong>in</strong> f<strong>in</strong>ite, then there is an upper bound to This implies that choos<strong>in</strong>g<br />

→∞means that ()−(0)<br />

<br />

→ 0 In both cases the left-hand side of the equation tends to zero giv<strong>in</strong>g the<br />

Virial theorem<br />

* +<br />

h i = − 1 X<br />

F · r (2.86)<br />

2<br />

<br />

The right-hand side of this equation is called the Virial of the system. For a s<strong>in</strong>gle particle subject to a<br />

conservative central force F = −∇ the Virial theorem equals<br />

h i = 1 2 h∇ · ri = 1 ¿<br />

À<br />

(2.87)<br />

2 <br />

If the potential is of the form = +1 that is, = −( +1) ,then <br />

<br />

=( +1). Thus for a s<strong>in</strong>gle<br />

particle <strong>in</strong> a central potential = +1 the Virial theorem reduces to<br />

h i = +1 hi (2.88)<br />

2<br />

The follow<strong>in</strong>g two special cases are of considerable importance <strong>in</strong> physics.<br />

Hooke’s Law: Notethatforal<strong>in</strong>earrestor<strong>in</strong>gforce =1then<br />

h i =+hi ( =1)<br />

You may be familiar with this fact for simple harmonic motion where the average k<strong>in</strong>etic and potential<br />

energies are the same and both equal half of the total energy.


2.11. VIRIAL THEOREM 23<br />

Inverse-square law:<br />

Theother<strong>in</strong>terest<strong>in</strong>gcaseisforthe<strong>in</strong>versesquarelaw = −2 where<br />

h i = − 1 hi ( = −2)<br />

2<br />

The Virial theorem is useful for solv<strong>in</strong>g problems <strong>in</strong> that know<strong>in</strong>g the exponent of the field makes it<br />

possible to write down directly the average total energy <strong>in</strong> the field. For example, for = −2<br />

hi = h i + hi = − 1 2 hi + hi = 1 hi (2.89)<br />

2<br />

This occurs for the Bohr model of the hydrogen atom where the k<strong>in</strong>etic energy of the bound electron is half<br />

of the potential energy. The same result occurs for planetary motion <strong>in</strong> the solar system.<br />

2.5 Example: The ideal gas law<br />

The Virial theorem deals with average properties and has applications to statistical mechanics. Consider<br />

an ideal gas. Accord<strong>in</strong>g to the Equipartition theorem the average k<strong>in</strong>etic energy per atom <strong>in</strong> an ideal gas is<br />

3<br />

2 where is the absolute temperature and is the Boltzmann constant. Thus the average total k<strong>in</strong>etic<br />

energy for atoms is hi = 3 2 . The right-hand side of the Virial theorem conta<strong>in</strong>s the force .For<br />

an ideal gas it is assumed that there are no <strong>in</strong>teraction forces between atoms, that is the only force is the<br />

force of constra<strong>in</strong>t of the walls of the pressure vessel. The pressure is force per unit area and thus the<br />

<strong>in</strong>stantaneous force on an area of wall is F = −ˆn where ˆ designates the unit vector normal to<br />

the surface. Thus the right-hand side of the Virial theorem is<br />

* +<br />

− 1 X<br />

F · r = Z<br />

ˆn · r <br />

2<br />

2<br />

<br />

Use of the divergence theorem thus gives that R ˆn·r = R ∇ · r =3 R =3 Thus the Virial theorem<br />

leads to the ideal gas law, that is<br />

= <br />

2.6 Example: The mass of galaxies<br />

The Virial theorem can be used to make a crude estimate of the mass of a cluster of galaxies. Assum<strong>in</strong>g a<br />

spherically-symmetric cluster of galaxies,eachofmass then the total mass of the cluster is = .<br />

A crude estimate of the cluster potential energy is<br />

hi ≈ 2<br />

<br />

()<br />

where is the radius of a cluster. The average k<strong>in</strong>etic energy per galaxy is 1 2 hi2 where hi 2 is the average<br />

square of the galaxy velocities with respect to the center of mass of the cluster. Thus the total k<strong>in</strong>etic energy<br />

of the cluster is<br />

hi ≈ hi2 = hi2<br />

()<br />

2 2<br />

The Virial theorem tells us that a central force hav<strong>in</strong>g a radial dependence of the form ∝ gives hi =<br />

+1<br />

2<br />

hi. For the <strong>in</strong>verse-square gravitational force then<br />

hi = − 1 2 hi<br />

()<br />

Thus equations and give an estimate of the total mass of the cluster to be<br />

≈ hi2<br />

<br />

This estimate is larger than the value estimated from the lum<strong>in</strong>osity of the cluster imply<strong>in</strong>g a large amount<br />

of “dark matter” must exist <strong>in</strong> galaxies which rema<strong>in</strong>s an open question <strong>in</strong> physics.


24 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

2.12 Applications of Newton’s equations of motion<br />

Newton’s equation of motion can be written <strong>in</strong> the form<br />

F = p<br />

= v = r<br />

2 2 (2.90)<br />

A description of the motion of a particle requires a solution of this second-order differential equation of<br />

motion. This equation of motion may be <strong>in</strong>tegrated to f<strong>in</strong>d r() and v() if the <strong>in</strong>itial conditions and<br />

the force field F() are known. Solution of the equation of motion can be complicated for many practical<br />

examples, but there are various approaches to simplify the solution. It is of value to learn efficient approaches<br />

to solv<strong>in</strong>g problems.<br />

The follow<strong>in</strong>g sequence is recommended<br />

a) Make a vector diagram of the problem <strong>in</strong>dicat<strong>in</strong>g forces, velocities, etc.<br />

b) Write down the known quantities.<br />

c) Before try<strong>in</strong>g to solve the equation of motion directly, look to see if a basic conservation law applies.<br />

That is, check if any of the three first-order <strong>in</strong>tegrals, can be used to simplify the solution. The use of<br />

conservation of energy or conservation of momentum can greatly simplify solv<strong>in</strong>g problems.<br />

The follow<strong>in</strong>g examples show the solution of typical types of problem encountered us<strong>in</strong>g Newtonian<br />

mechanics.<br />

2.12.1 Constant force problems<br />

Problems hav<strong>in</strong>g a constant force imply constant acceleration. The classic example is a block slid<strong>in</strong>g on an<br />

<strong>in</strong>cl<strong>in</strong>ed plane, where the block of mass is acted upon by both gravity and friction. The net force F is<br />

given by the vector sum of the gravitational force F ,normalforceN and frictional force f .<br />

Tak<strong>in</strong>g components perpendicular to the <strong>in</strong>cl<strong>in</strong>ed plane <strong>in</strong> the direction<br />

That is, s<strong>in</strong>ce = <br />

F = F + N + f = a (2.91)<br />

− cos + =0 (2.92)<br />

= cos (2.93)<br />

Similarly, tak<strong>in</strong>g components along the <strong>in</strong>cl<strong>in</strong>ed plane <strong>in</strong> the direction<br />

s<strong>in</strong> − = 2 <br />

2 (2.94)<br />

Us<strong>in</strong>g the concept of coefficient of friction <br />

f f<br />

N<br />

y<br />

Thus the equation of motion can be written as<br />

= (2.95)<br />

(s<strong>in</strong> − cos ) = 2 <br />

2 (2.96)<br />

The block accelerates if s<strong>in</strong> cos that is, tan The<br />

acceleration is constant if and are constant, that is<br />

Fg<br />

Figure 2.3: Block on an <strong>in</strong>cl<strong>in</strong>ed plane<br />

2 <br />

= (s<strong>in</strong> − cos ) (2.97)<br />

2 Remember that if the block is stationary, the friction coefficient balances such that (s<strong>in</strong> − cos ) =0<br />

that is, tan = . However, there is a maximum static friction coefficient beyond which the block starts<br />

slid<strong>in</strong>g. The k<strong>in</strong>etic coefficient of friction is applicable for slid<strong>in</strong>g friction and usually <br />

Another example of constant force and acceleration is motion of objects free fall<strong>in</strong>g <strong>in</strong> a uniform gravitational<br />

field when air drag is neglected. Then one obta<strong>in</strong>s the simple relations such as = + , etc.<br />

x


2.12. APPLICATIONS OF NEWTON’S EQUATIONS OF MOTION 25<br />

2.12.2 L<strong>in</strong>ear Restor<strong>in</strong>g Force<br />

An important class of problems <strong>in</strong>volve a l<strong>in</strong>ear restor<strong>in</strong>g force, that is, they obey Hooke’s law. The equation<br />

of motion for this case is<br />

() =− = ¨ (2.98)<br />

It is usual to def<strong>in</strong>e<br />

2 0 ≡ <br />

(2.99)<br />

Then the equation of motion then can be written as<br />

¨ + 2 0 =0 (2.100)<br />

which is the equation of the harmonic oscillator. Examples are small oscillations of a mass on a spr<strong>in</strong>g,<br />

vibrations of a stretched piano str<strong>in</strong>g, etc.<br />

The solution of this second order equation is<br />

() = s<strong>in</strong> ( 0 − ) (2.101)<br />

This is the well known s<strong>in</strong>usoidal behavior of the displacement for the simple harmonic oscillator.<br />

angular frequency 0 is<br />

The<br />

r<br />

<br />

0 =<br />

<br />

(2.102)<br />

Note that this l<strong>in</strong>ear system has no dissipative forces, thus the total energy is a constant of motion as<br />

discussed previously. That is, it is a conservative system with a total energy given by<br />

1<br />

2 ˙2 + 1 2 2 = (2.103)<br />

The first term is the k<strong>in</strong>etic energy and the second term is the potential energy. The Virial theorem gives<br />

that for the l<strong>in</strong>ear restor<strong>in</strong>g force the average k<strong>in</strong>etic energy equals the average potential energy.<br />

2.12.3 Position-dependent conservative forces<br />

The l<strong>in</strong>ear restor<strong>in</strong>g force is an example of a conservative field. The total energy is conserved, and if the<br />

field is time <strong>in</strong>dependent, then the conservative forces are a function only of position. The easiest way to<br />

solve such problems is to use the concept of potential energy illustrated <strong>in</strong> Figure 24.<br />

2 − 1 = −<br />

Z 2<br />

1<br />

F · r (2.104)<br />

Consider a conservative force <strong>in</strong> one dimension. S<strong>in</strong>ce it was shown that the total energy = + is<br />

conserved for a conservative field, then<br />

= + = 1 2 2 + () (2.105)<br />

Therefore:<br />

Integration of this gives<br />

= <br />

r<br />

2<br />

= ± [ − ()] (2.106)<br />

<br />

− 0 =<br />

Z <br />

0<br />

±<br />

q<br />

(2.107)<br />

2<br />

[ − ()]<br />

where = 0 when = 0 Know<strong>in</strong>g () it is possible to solve this equation as a function of time.


26 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

It is possible to understand the general features of the<br />

solution just from <strong>in</strong>spection of the function () For example,<br />

as shown <strong>in</strong> figure 24 the motion for energy 1<br />

is periodic between the turn<strong>in</strong>g po<strong>in</strong>ts and S<strong>in</strong>ce<br />

the potential energy curve is approximately parabolic between<br />

these limits the motion will exhibit simple harmonic<br />

motion. For 0 the turn<strong>in</strong>g po<strong>in</strong>t coalesce to 0 that is<br />

U(x)<br />

E 4<br />

E 3<br />

E 2<br />

E 1<br />

E 0<br />

x x x x<br />

g<br />

x<br />

a o b<br />

x<br />

c d e f<br />

Figure 2.4: One-dimensional potential ()<br />

there is no motion. For total energy 2 the motion is<br />

periodic <strong>in</strong> two <strong>in</strong>dependent regimes, ≤ ≤ and<br />

≤ ≤ <strong>Classical</strong>ly the particle cannot jump from<br />

one pocket to the other. The motion for the particle with<br />

total energy 3 is that it moves freely from <strong>in</strong>f<strong>in</strong>ity, stops<br />

and rebounds at = andthenreturnsto<strong>in</strong>f<strong>in</strong>ity. That<br />

is the particle bounces off the potential at For energy<br />

4 the particle moves freely and is unbounded. For all<br />

these cases, the actual velocity is given by the above relation<br />

for () Thus the k<strong>in</strong>etic energy is largest where<br />

the potential is deepest. An example would be motion of<br />

a roller coaster car.<br />

Position-dependent forces are encountered extensively<br />

<strong>in</strong> classical mechanics. Examples are the many manifestations<br />

of motion <strong>in</strong> gravitational fields, such as <strong>in</strong>terplanetary<br />

probes, a roller coaster, and automobile suspension systems. The l<strong>in</strong>ear restor<strong>in</strong>g force is an especially<br />

simple example of a position-dependent force while the most frequently encountered conservative potentials<br />

are <strong>in</strong> electrostatics and gravitation for which the potentials are;<br />

() = 1 1 2<br />

4 0 12<br />

2 (Electrostatic potential energy)<br />

() =− 1 2<br />

12<br />

2 (Gravitational potential energy)<br />

Know<strong>in</strong>g () it is possible to solve the equation of motion as a function of time.<br />

2.7 Example: Diatomic molecule<br />

An example of a conservative field is a vibrat<strong>in</strong>g diatomic molecule which has a potential energy dependence<br />

with separation distance that is described approximately by the Morse function<br />

h<br />

i<br />

() = 0 1 − − (− 0 ) 2<br />

− 0<br />

where 0 0 and are parameters chosen to best describe the particular pair of atoms. The restor<strong>in</strong>g force<br />

is given by<br />

() =− () =2 i<br />

0<br />

h1 − − (− 0 )<br />

<br />

ih − (− 0 )<br />

<br />

<br />

This has a m<strong>in</strong>imum value of ( 0 )= 0 at = 0 <br />

Note that for small amplitude oscillations, where<br />

( − 0 ) <br />

the exponential term <strong>in</strong> the potential function can be expanded<br />

to give<br />

∙<br />

() ≈ 0 1 − (1 −− ( − ¸2<br />

0)<br />

) − 0 ≈ 0<br />

<br />

2 (− 0) 2 − 0<br />

This gives a restor<strong>in</strong>g force<br />

() =− () = −2 0<br />

( − 0)<br />

That is, for small amplitudes the restor<strong>in</strong>g force is l<strong>in</strong>ear.<br />

U(x)<br />

U o<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

-0.2<br />

-0.4<br />

-0.6<br />

-0.8<br />

-1.0<br />

x<br />

1 2 3 4 5<br />

Potential energy function () 0 versus <br />

for the diatomic molecule.<br />

x<br />

x<br />

x


2.12. APPLICATIONS OF NEWTON’S EQUATIONS OF MOTION 27<br />

2.12.4 Constra<strong>in</strong>ed motion<br />

A frequently encountered problem <strong>in</strong>volv<strong>in</strong>g position dependent forces, is when the motion is constra<strong>in</strong>ed to<br />

follow a certa<strong>in</strong> trajectory. Forces of constra<strong>in</strong>t must exist to constra<strong>in</strong> the motion to a specific trajectory.<br />

Examples are, the roller coaster, a roll<strong>in</strong>g ball on an undulat<strong>in</strong>g surface, or a downhill skier, where the<br />

motion is constra<strong>in</strong>ed to follow the surface or track contours. The potential energy can be evaluated at all<br />

positions along the constra<strong>in</strong>ed trajectory for conservative forces such as gravity. However, the additional<br />

forces of constra<strong>in</strong>t that must exist to constra<strong>in</strong> the motion, can be complicated and depend on the motion.<br />

For example, the roller coaster must always balance the gravitational and centripetal forces. Fortunately<br />

forces of constra<strong>in</strong>t F often are normal to the direction of motion and thus do not contribute to the total<br />

mechanical energy s<strong>in</strong>ce then the work done F · l is zero. Magnetic forces F =v × B exhibit this feature<br />

of hav<strong>in</strong>g the force normal to the motion.<br />

Solution of constra<strong>in</strong>ed problems is greatly simplifiediftheotherforcesareconservativeandtheforces<br />

of constra<strong>in</strong>t are normal to the motion, s<strong>in</strong>ce then energy conservation can be used.<br />

2.8 Example: Roller coaster<br />

Consider motion of a roller coaster shown <strong>in</strong> the<br />

adjacent figure. This system is conservative if the friction<br />

and air drag are neglected and then the forces of<br />

constra<strong>in</strong>t are normal to the direction of motion.<br />

The k<strong>in</strong>etic energy at any position is just given by<br />

energy conservation and the fact that<br />

= + <br />

where depends on the height of the track at any the<br />

given location. The k<strong>in</strong>etic energy is greatest when the<br />

potential energy is lowest. The forces of constra<strong>in</strong>t<br />

can be deduced if the velocity of motion on the track<br />

is known. Assum<strong>in</strong>g that the motion is conf<strong>in</strong>edtoa<br />

vertical plane, then one has a centripetal force of constra<strong>in</strong>t<br />

2<br />

<br />

normal to the track <strong>in</strong>wards towards the<br />

center of the radius of curvature , plus the gravitation<br />

force downwards of <br />

The constra<strong>in</strong>t force is 2 <br />

<br />

− upwards at the<br />

top of the loop, while it is 2 <br />

<br />

+ downwards at<br />

the bottom of the loop. To ensure that the car and<br />

occupants do not leave the required trajectory, the force<br />

upwards at the top of the loop has to be positive, that<br />

is, 2 ≥ . The velocity at the bottom of the loop<br />

is given by 1 2 2 = 1 2 2 +2 assum<strong>in</strong>g that the<br />

track has a constant radius of curvature . That is;<br />

at a m<strong>in</strong>imum 2 = +4 =5 Therefore the<br />

occupants now will feel an acceleration downwards of<br />

Roller coaster (CCO Public Doma<strong>in</strong>)<br />

at least 2 <br />

<br />

+ =6 at the bottom of the loop The<br />

first roller coaster was built with such a constant radius of curvature but an acceleration of 6 was too much<br />

for the average passenger. Therefore roller coasters are designed such that the radius of curvature is much<br />

larger at the bottom of the loop, as illustrated, <strong>in</strong> order to ma<strong>in</strong>ta<strong>in</strong> sufficiently low loads and also ensure<br />

that the required constra<strong>in</strong>t forces exist.<br />

Note that the m<strong>in</strong>imum velocity at the top of the loop, , implies that if the cart starts from rest it must<br />

startataheight > 2<br />

above the top of the loop if friction is negligible. Note that the solution for the roll<strong>in</strong>g<br />

ball on such a roller coaster differs from that for a slid<strong>in</strong>g object s<strong>in</strong>ce one must <strong>in</strong>clude the rotational energy<br />

of the ball as well as the l<strong>in</strong>ear velocity.<br />

Loop<strong>in</strong>g the loop <strong>in</strong> a sailplane <strong>in</strong>volves the same physics mak<strong>in</strong>g it necessary to vary the elevator control<br />

to vary the radius of curvature throughout the loop to m<strong>in</strong>imize the maximum load.


28 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

2.12.5 Velocity Dependent Forces<br />

Velocity dependent forces are encountered frequently <strong>in</strong> practical problems. For example, motion of an<br />

object <strong>in</strong> a fluid, such as air, where viscous forces retard the motion. In general the retard<strong>in</strong>g force has a<br />

complicated dependence on velocity. A quadrative-velocity drag force <strong>in</strong> air often can be expressed <strong>in</strong> the<br />

form,<br />

F () =− 1 2 2 bv (2.108)<br />

where is a dimensionless drag coefficient, is the density of air, is the cross sectional area perpendicular<br />

to the direction of motion, and is the velocity. Modern automobiles have drag coefficients as low as 03. As<br />

described<strong>in</strong>chapter16, the drag coefficient depends on the Reynold’s number which relates the <strong>in</strong>ertial to<br />

viscous drag forces. Small sized objects at low velocity, such as light ra<strong>in</strong>drops, have low Reynold’s numbers<br />

for which is roughly proportional to −1 lead<strong>in</strong>g to a l<strong>in</strong>ear dependence of the drag force on velocity, i.e.<br />

() ∝ . Larger objects mov<strong>in</strong>g at higher velocities, suchasacarorsky-diver,havehigherReynold’s<br />

numbers for which is roughly <strong>in</strong>dependent of velocity lead<strong>in</strong>g to a drag force () ∝ 2 .Thisdragforce<br />

always po<strong>in</strong>ts <strong>in</strong> the opposite direction to the unit velocity vector. Approximately for air<br />

F () =− ¡ 1 + 2 2¢ bv (2.109)<br />

where for spherical objects of diameter , 1 ≈ 155×10 −4 and 2 ≈ 022 2 <strong>in</strong> MKS units. Fortunately, the<br />

equation of motion usually can be <strong>in</strong>tegrated when the retard<strong>in</strong>g force has a simple power law dependence.<br />

As an example, consider free fall <strong>in</strong> the Earth’s gravitational field.<br />

2.9 Example: Vertical fall <strong>in</strong> the earth’s gravitational field.<br />

L<strong>in</strong>ear regime<br />

1 2 <br />

For small objects at low-velocity, i.e. low Reynold’s number, the drag approximately has a l<strong>in</strong>ear dependence<br />

on velocity. Then the equation of motion is<br />

Separate the variables and <strong>in</strong>tegrate<br />

That is<br />

=<br />

Z <br />

0<br />

− − 1 = <br />

<br />

µ <br />

<br />

+<br />

− − 1 = − 1 <br />

ln<br />

1 + 1 0<br />

= − µ <br />

+ + 0 − 1<br />

<br />

1 1<br />

Note that for À 1<br />

the velocity approaches a term<strong>in</strong>al velocity of ∞ = − <br />

1<br />

Thecharacteristictime<br />

constant is = 1<br />

= ∞ Note that if 0 =0 then<br />

´<br />

= ∞<br />

³1 − − <br />

For the case of small ra<strong>in</strong>drops with =05 then ∞ =8 (18) and time constant =08sec<br />

Note that <strong>in</strong> the absence of air drag, these ra<strong>in</strong> drops fall<strong>in</strong>g from 2000 would atta<strong>in</strong> a velocity of over<br />

400 m.p.h. It is fortunate that the drag reduces the speed of ra<strong>in</strong> drops to non-damag<strong>in</strong>g values. Note that<br />

the above relation would predict high velocities for hail. Fortunately, the drag <strong>in</strong>creases quadratically at the<br />

higher velocities atta<strong>in</strong>ed by large ra<strong>in</strong> drops or hail, and this limits the term<strong>in</strong>al velocity to moderate values.<br />

For the United States these velocities still are sufficienttodoconsiderablecropdamage<strong>in</strong>themid-west.<br />

Quadratic regime 2 1<br />

For larger objects at higher velocities, i.e. high Reynold’s number, the drag depends on the square of the<br />

velocity mak<strong>in</strong>g it necessary to differentiate between objects ris<strong>in</strong>g and fall<strong>in</strong>g. The equation of motion is<br />

− ± 2 2 =


2.12. APPLICATIONS OF NEWTON’S EQUATIONS OF MOTION 29<br />

where the positive sign is for fall<strong>in</strong>g objects and negative sign for ris<strong>in</strong>g objects. Integrat<strong>in</strong>g the equation of<br />

motion for fall<strong>in</strong>g gives<br />

Z <br />

µ<br />

<br />

<br />

=<br />

− + 2 2 = tanh −1 0<br />

− tanh −1 <br />

∞ ∞<br />

q<br />

where = <br />

and 2 ∞ =<br />

velocity gives<br />

0<br />

q<br />

<br />

2<br />

That is, = ∞<br />

<br />

For the case of a fall<strong>in</strong>g object with 0 =0 solv<strong>in</strong>g for<br />

= ∞ tanh <br />

As an example, a 06 basket ball with =025 will have ∞ =20 ( 43 m.p.h.) and =21.<br />

Consider President George H.W. Bush skydiv<strong>in</strong>g. Assume his mass is 70kg and assume an equivalent<br />

spherical shape of the former President to have a diameter of =1. This gives that ∞ =56<br />

( 120) and =56. When Bush senior opens his 8 diameter parachute his term<strong>in</strong>al velocity is<br />

estimated to decrease to 7 ( 15 ) which is close to the value for a typical ( 8) diameter emergency<br />

parachute which has a measured term<strong>in</strong>al velocity of 11 <strong>in</strong> spite of air leakage through the central vent<br />

needed to stabilize the parachute motion.<br />

2.10 Example: Projectile motion <strong>in</strong> air<br />

Consider a projectile <strong>in</strong>itially at = =0at =0,thatisfired at an <strong>in</strong>itial velocity v 0 at an angle<br />

to the horizontal. In order to understand the general features of the solution, assume that the drag is<br />

proportional to velocity. This is <strong>in</strong>correct for typical projectile velocities, but simplifies the mathematics. The<br />

equations of motion can be expressed as<br />

¨ = − ˙<br />

¨ = − ˙ − <br />

where is the coefficient for air drag. Take the <strong>in</strong>itial conditions at =0to be = =0 ˙ = cos <br />

˙ = s<strong>in</strong> <br />

Solv<strong>in</strong>g <strong>in</strong> the x coord<strong>in</strong>ate,<br />

˙<br />

= − ˙<br />

<br />

Therefore<br />

˙ = cos −<br />

That is, the velocity decays to zero with a time constant = 1 .<br />

Integration of the velocity equation gives<br />

= <br />

<br />

¡<br />

1 − <br />

− ¢<br />

Note that this implies that the body approaches a value of = <br />

<br />

as →∞<br />

The trajectory of an object is distorted from the parabolic shape, that occurs for =0 due to the rapid<br />

drop <strong>in</strong> range as the drag coefficient <strong>in</strong>creases. For realistic cases it is necessary to use a computer to solve<br />

this numerically.<br />

2.12.6 Systems with Variable Mass<br />

Classic examples of systems with variable mass are the rocket, a fall<strong>in</strong>g cha<strong>in</strong>, and nuclear fission. Consider<br />

the problem of vertical rocket motion <strong>in</strong> a gravitational field us<strong>in</strong>g Newtonian mechanics. When there is a<br />

vertical gravitational external field, the vertical momentum is not conserved due to both gravity and the<br />

ejection of rocket propellant. In a time the rocket ejects propellant vertically with exhaust velocity<br />

relative to the rocket of . Thus the momentum imparted to this propellant is<br />

= − (2.110)<br />

Therefore the rocket is given an equal and opposite <strong>in</strong>crease <strong>in</strong> momentum <br />

=+ (2.111)


30 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

In the time <strong>in</strong>terval the net change <strong>in</strong> the l<strong>in</strong>ear momentum of the rocket plus fuel system is given by<br />

=( − )( + )+ ( − ) − = − (2.112)<br />

Therateofchangeofthel<strong>in</strong>earmomentumthusequals<br />

= <br />

= − <br />

<br />

(2.113)<br />

Consider the problem for the special case of vertical ascent of the rocket aga<strong>in</strong>st the external gravitational<br />

force = −. Then<br />

This can be rewritten as<br />

− + <br />

<br />

= <br />

<br />

− + ˙ = ˙<br />

The second term comes from the variable rocket mass where<br />

the loss of mass of the rocket equals the mass of the ejected<br />

propellant. Assum<strong>in</strong>g a constant fuel burn ˙ = then<br />

˙ = − ˙ = − (2.115)<br />

y<br />

v<br />

(2.114)<br />

where 0 Then the equation becomes<br />

=<br />

³− + ´<br />

(2.116)<br />

<br />

m<br />

g<br />

S<strong>in</strong>ce<br />

then<br />

<br />

<br />

− <br />

= − (2.117)<br />

= (2.118)<br />

Earth<br />

dm’<br />

u<br />

Insert<strong>in</strong>g this <strong>in</strong> the above equation gives<br />

³ <br />

=<br />

´<br />

− (2.119)<br />

Figure 2.5: Vertical motion of a rocket <strong>in</strong> a<br />

gravitational field<br />

Integration gives<br />

But the change <strong>in</strong> mass is given by<br />

That is<br />

= − ( 0 − )+ ln<br />

Z <br />

0<br />

= −<br />

Z <br />

0<br />

³ 0<br />

´<br />

<br />

(2.120)<br />

(2.121)<br />

0 − = (2.122)<br />

Thus<br />

³ 0<br />

´<br />

= − + ln<br />

(2.123)<br />

<br />

Note that once the propellant is exhausted the rocket will cont<strong>in</strong>ue to fly upwards as it decelerates <strong>in</strong><br />

the gravitational field. You can easily calculate the maximum height. Note that this formula assumes that<br />

the acceleration due to gravity is constant whereas for large heights above the Earth it is necessary to use<br />

the true gravitational force − <br />

<br />

where isthedistancefromthecenteroftheearth. Inrealsituations<br />

2<br />

it is necessary to <strong>in</strong>clude air drag which requires a computer to numerically solve the equations of motion.<br />

The highest rocket velocity is atta<strong>in</strong>ed by maximiz<strong>in</strong>g the exhaust velocity and the ratio of <strong>in</strong>itial to f<strong>in</strong>al<br />

mass. Because the term<strong>in</strong>al velocity is limited by the mass ratio, eng<strong>in</strong>eers construct multistage rockets that<br />

jettison the spent fuel conta<strong>in</strong>ers and rockets. The variational-pr<strong>in</strong>ciple approach applied to variable mass<br />

problems is discussed <strong>in</strong> chapter 87


2.12. APPLICATIONS OF NEWTON’S EQUATIONS OF MOTION 31<br />

2.12.7 Rigid-body rotation about a body-fixedrotationaxis<br />

The most general case of rigid-body rotation <strong>in</strong>volves rotation about some body-fixed po<strong>in</strong>t with the orientation<br />

of the rotation axis undef<strong>in</strong>ed. For example, an object sp<strong>in</strong>n<strong>in</strong>g <strong>in</strong> space will rotate about the center<br />

of mass with the rotation axis hav<strong>in</strong>g any orientation. Another example is a child’s sp<strong>in</strong>n<strong>in</strong>g top which sp<strong>in</strong>s<br />

with arbitrary orientation of the axis of rotation about the po<strong>in</strong>ted end which touches the ground about a<br />

static location. Such rotation about a body-fixed po<strong>in</strong>t is complicated and will be discussed <strong>in</strong> chapter 13.<br />

Rigid-body rotation is easier to handle if the orientation of the axis of rotation is fixed with respect to the<br />

rigid body. An example of such motion is a h<strong>in</strong>ged door.<br />

For a rigid body rotat<strong>in</strong>g with angular velocity the total angular momentum L is given by<br />

L =<br />

X<br />

L =<br />

<br />

X<br />

r × p (2.124)<br />

<br />

For rotation equation appendix 29 gives<br />

thus the angular momentum can be written as<br />

v = ω × r <br />

(294)<br />

L =<br />

X<br />

r × p =<br />

<br />

X<br />

r × ω × r (2.125)<br />

<br />

The vector triple product can be simplified us<strong>in</strong>g the vector identity equation 24 giv<strong>in</strong>g<br />

L =<br />

X £¡<br />

2 ¢ ¤<br />

ω − (r · ω) r <br />

<br />

(2.126)<br />

Rigid-body rotation about a body-fixed symmetry axis<br />

The simplest case for rigid-body rotation is when the body has a symmetry axis with the angular velocity ω<br />

parallel to this body-fixed symmetry axis. For this case then r can be taken perpendicular to ω for which<br />

the second term <strong>in</strong> equation 2126, i.e. (r · ω) =0,thus<br />

L =<br />

X<br />

<br />

¡<br />

<br />

2 ¢<br />

ω (r perpendicular to ω)<br />

The moment of <strong>in</strong>ertia about the symmetry axis is def<strong>in</strong>ed as<br />

=<br />

X<br />

2 (2.127)<br />

<br />

where is the perpendicular distance from the axis of rotation to the body, For a cont<strong>in</strong>uous body the<br />

moment of <strong>in</strong>ertia can be generalized to an <strong>in</strong>tegral over the mass density of the body<br />

Z<br />

= 2 (2.128)<br />

where is perpendicular to the rotation axis. The def<strong>in</strong>ition of the moment of <strong>in</strong>ertia allows rewrit<strong>in</strong>g the<br />

angular momentum about a symmetry axis L <strong>in</strong> the form<br />

L = ω (2.129)<br />

where the moment of <strong>in</strong>ertia is taken about the symmetry axis and assum<strong>in</strong>g that the angular velocity<br />

of rotation vector is parallel to the symmetry axis.


32 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

Rigid-body rotation about a non-symmetric body-fixed axis<br />

In general the fixed axis of rotation is not aligned with a symmetry axis of the body, or the body does not<br />

have a symmetry axis, both of which complicate the problem.<br />

For illustration consider that the rigid body comprises a system of masses located at positions r <br />

with the rigid body rotat<strong>in</strong>g about the axis with angular velocity ω That is,<br />

ω = ẑ (2.130)<br />

In cartesian coord<strong>in</strong>ates the fixed-frame vector for particle is<br />

r =( ) (2.131)<br />

us<strong>in</strong>g these <strong>in</strong> the cross product (294) gives<br />

⎛<br />

v = ω × r = ⎝ − <br />

<br />

0<br />

⎞<br />

⎠ (2.132)<br />

which is written as a column vector for clarity. Insert<strong>in</strong>g v <strong>in</strong> the cross-product r ×v gives the components<br />

of the angular momentum to be<br />

⎛<br />

X<br />

X<br />

L = r × v = <br />

⎝<br />

− ⎞<br />

<br />

− <br />

⎠<br />

<br />

2 + 2 <br />

That is, the components of the angular momentum are<br />

à <br />

!<br />

X<br />

= − ≡ (2.133)<br />

<br />

à <br />

!<br />

X<br />

= − ≡ <br />

=<br />

<br />

Ã<br />

X <br />

£<br />

<br />

2<br />

+ 2 ¤ ! ≡ <br />

<br />

Note that the perpendicular distance from the axis<br />

<strong>in</strong> cyl<strong>in</strong>drical coord<strong>in</strong>ates is = p 2 + 2 <br />

thus the angular<br />

momentum about the axis can be written<br />

as<br />

à <br />

!<br />

X<br />

= 2 = (2.134)<br />

<br />

where (2134) gives the elementary formula for the moment<br />

of <strong>in</strong>ertia = about the axis given earlier<br />

<strong>in</strong> (2129).<br />

The surpris<strong>in</strong>g result is that and are non-zero<br />

imply<strong>in</strong>g that the total angular momentum vector L is<br />

<strong>in</strong> general not parallel with ω This can be understood<br />

by consider<strong>in</strong>g the s<strong>in</strong>gle body shown <strong>in</strong> figure 26.<br />

When the body is <strong>in</strong> the plane then = 0 and<br />

=0 Thus the angular momentum vector L has a<br />

component along the − direction as shown which is<br />

not parallel with ω and, s<strong>in</strong>ce the vectors ω L r are<br />

L<br />

x<br />

Figure 2.6: A rigid rotat<strong>in</strong>g body compris<strong>in</strong>g a s<strong>in</strong>gle<br />

mass attached by a massless rod at a fixed<br />

angle shown at the <strong>in</strong>stant when happens to<br />

lie <strong>in</strong> the plane. As the body rotates about<br />

the − axis the mass has a velocity and momentum<br />

<strong>in</strong>to the page (the negative direction).<br />

Therefore the angular momentum L = r × p is <strong>in</strong><br />

the direction shown which is not parallel to the<br />

angular velocity <br />

coplanar, then L must sweep around the rotation axis ω to rema<strong>in</strong> coplanar with the body as it rotates<br />

about the axis. Instantaneously the velocity of the body v is <strong>in</strong>to the plane of the paper and, s<strong>in</strong>ce<br />

z<br />

O<br />

r<br />

y


2.12. APPLICATIONS OF NEWTON’S EQUATIONS OF MOTION 33<br />

L = r × v then L is at an angle (90 ◦ − ) to the axis. This implies that a torque must be applied<br />

to rotate the angular momentum vector. This expla<strong>in</strong>s why your automobile shakes if the rotation axis and<br />

symmetry axis are not parallel for one wheel.<br />

The first two moments <strong>in</strong> (2133) are called products of <strong>in</strong>ertia of the body designated by the pair of<br />

axes <strong>in</strong>volved. Therefore, to avoid confusion, it is necessary to def<strong>in</strong>e the diagonal moment, which is called<br />

the moment of <strong>in</strong>ertia, by two subscripts as Thus <strong>in</strong> general, a body can have three moments of <strong>in</strong>ertia<br />

about the three axes plus three products of <strong>in</strong>ertia. This group of moments comprise the <strong>in</strong>ertia tensor<br />

which will be discussed further <strong>in</strong> chapter 13. Ifabodyhasanaxisofsymmetryalongthe axis then the<br />

summations will give = =0while will be unchanged. That is, for rotation about a symmetry<br />

axis the angular momentum and rotation axes are parallel. For any axis along which the angular momentum<br />

and angular velocity co<strong>in</strong>cide is called a pr<strong>in</strong>cipal axis of the body.<br />

2.11 Example: Moment of <strong>in</strong>ertia of a th<strong>in</strong> door<br />

Consider that the door has width and height and assume the door thickness is negligible with areal<br />

density 2. Assume that the door is h<strong>in</strong>ged about the axis. The mass of a surface element of<br />

dimension at a distance from the rotation axis is = thus the mass of the complete door<br />

is = The moment of <strong>in</strong>ertia about the axis is given by<br />

=<br />

Z <br />

=0<br />

Z <br />

=0<br />

2 = 1 3 3 = 1 3 2<br />

2.12 Example: Merry-go-round<br />

A child of mass jumps onto the outside edge of a circular merry-go-round of moment of <strong>in</strong>ertia , and<br />

radius and <strong>in</strong>itial angular velocity 0 What is the f<strong>in</strong>al angular velocity ?<br />

If the <strong>in</strong>itial angular momentum is 0 and, assum<strong>in</strong>g the child jumps with zero angular velocity, then the<br />

conservation of angular momentum implies that<br />

That is<br />

0 = <br />

0 = + <br />

0<br />

= <br />

( + 2 )<br />

<br />

= <br />

=<br />

0 0 + 2<br />

Note that this is true <strong>in</strong>dependent of the details of the acceleration of the <strong>in</strong>itially stationary child.<br />

2.13 Example: Cue pushes a billiard ball<br />

Cue push<strong>in</strong>g a billiard ball horizontally at the height<br />

of the centre of rotation of the ball.<br />

Consider a billiard ball of mass and radius <br />

is pushed by a cue <strong>in</strong> a direction that passes through<br />

the center of gravity such that the ball atta<strong>in</strong>s a velocity<br />

0 . The friction coefficient between the table and<br />

the ball is . How far does the ball move before the<br />

<strong>in</strong>itial slipp<strong>in</strong>g motion changes to pure roll<strong>in</strong>g motion?<br />

S<strong>in</strong>ce the direction of the cue force passes through<br />

the center of mass of the ball, it contributes zero<br />

torque to the ball. Thus the <strong>in</strong>itial angular momentum<br />

is zero at =0. The friction force po<strong>in</strong>ts opposite to the direction of motion and causes a torque <br />

about the center of mass <strong>in</strong> the direction ˆ.<br />

N = f · R =<br />

0


34 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

S<strong>in</strong>ce the moment of <strong>in</strong>ertia about the center of a uniform sphere is = 2 5 2 then the angular acceleration<br />

of the ball is<br />

˙ = = <br />

<br />

2<br />

= 5 <br />

()<br />

5 2 2 <br />

Moreover the frictional force causes a deceleration of the l<strong>in</strong>ear velocity of the center of mass of<br />

= − = −<br />

()<br />

Integrat<strong>in</strong>g from time zero to gives<br />

Z <br />

= ˙ = 5 <br />

0 2 <br />

The l<strong>in</strong>ear velocity of the center of mass at time is given by <strong>in</strong>tegration of equation <br />

=<br />

Z <br />

0<br />

= 0 − <br />

The billiard ball stops slid<strong>in</strong>g and only rolls when = , thatis,when<br />

5 <br />

2 = 0 − <br />

That is, when<br />

Thus the ball slips for a distance<br />

0<br />

= 2 7 <br />

=<br />

Z <br />

0<br />

= 0 − 2 <br />

2<br />

= 12 0<br />

2<br />

49 <br />

Note that if the ball is pushed at a distance above the center of mass, besides the l<strong>in</strong>ear velocity there<br />

is an <strong>in</strong>itial angular momentum of<br />

= 0<br />

2<br />

= 5 0 <br />

5 2 2 2<br />

For the special case = 2 5 the ball immediately assumes a pure non-slipp<strong>in</strong>g roll. For 2 5 one has<br />

0<br />

<br />

while 2 5 corresponds to 0<br />

<br />

. In the latter case the frictional force po<strong>in</strong>ts forward.<br />

2.12.8 Time dependent forces<br />

Many problems <strong>in</strong>volve action <strong>in</strong> the presence of a time dependent force. There are two extreme cases that<br />

are often encountered. One case is an impulsive force that acts for a very short time, for example, strik<strong>in</strong>g<br />

a ball with a bat, or the collision of two cars. The second case <strong>in</strong>volves an oscillatory time dependent force.<br />

The response to impulsive forces is discussed below whereas the response to oscillatory time-dependent forces<br />

is discussed <strong>in</strong> chapter 3.<br />

Translational impulsive forces<br />

An impulsive force acts for a very short time relative to the response time of the mechanical system be<strong>in</strong>g<br />

discussed. In pr<strong>in</strong>ciple the equation of motion can be solved if the complicated time dependence of the force,<br />

() is known. However, often it is possible to use the much simpler approach employ<strong>in</strong>g the concept of an<br />

impulse and the pr<strong>in</strong>ciple of the conservation of l<strong>in</strong>ear momentum.<br />

Def<strong>in</strong>e the l<strong>in</strong>ear impulse P to be the first-order time <strong>in</strong>tegral of the time-dependent force.<br />

Z<br />

P ≡ F() (2.135)


2.12. APPLICATIONS OF NEWTON’S EQUATIONS OF MOTION 35<br />

S<strong>in</strong>ce F() = p<br />

<br />

then equation 2135 gives that<br />

Z <br />

Z <br />

p<br />

P =<br />

0 0 0 = p = p() − p 0 = ∆p (2.136)<br />

0<br />

Thus the impulse P is an unambiguous quantity that equals the change <strong>in</strong> l<strong>in</strong>ear momentum of the object<br />

that has been struck which is <strong>in</strong>dependent of the details of the time dependence of the impulsive force.<br />

Computation of the spatial motion still requires knowledge of () s<strong>in</strong>ce the 2136 can be written as<br />

Z <br />

v() = 1 F( 0 ) 0 + v 0 (2.137)<br />

0<br />

Integration gives<br />

Z "<br />

Z #<br />

1 <br />

”<br />

r() − r 0 = v 0 +<br />

F( 0 ) 0 ” (2.138)<br />

0 0<br />

In general this is complicated. However, for the case of a constant force F() =F 0 this simplifies to the<br />

constant acceleration equation<br />

r() − r 0 = v 0 + 1 F 0<br />

2 2 (2.139)<br />

where the constant acceleration a = F 0<br />

.<br />

Angular impulsive torques<br />

Note that the pr<strong>in</strong>ciple of impulse also applies to angular motion. Def<strong>in</strong>e an impulsive torque T as the<br />

first-order time <strong>in</strong>tegral of the time-dependent torque.<br />

Z<br />

T ≡ N() (2.140)<br />

S<strong>in</strong>ce torque is related to the rate of change of angular momentum<br />

N() = L<br />

(2.141)<br />

<br />

then<br />

Z Z<br />

L<br />

<br />

T =<br />

0 0 0 = L = L() − L 0 = ∆L (2.142)<br />

0<br />

Thus the impulsive torque T equals the change <strong>in</strong> angular momentum ∆L of the struck body.<br />

2.14 Example: Center of percussion of a baseball bat<br />

When an impulsive force strikes a bat of mass at a distance<br />

s from the center of mass, then both the l<strong>in</strong>ear momentum<br />

of the center of mass, and angular momenta about the center<br />

of mass, of the bat are changed. Assume that the ball strikes<br />

the bat with an impulsive force = ∆ perpendicular to the<br />

symmetry axis of the bat at the strike po<strong>in</strong>t which is a distance<br />

from the center of mass of the bat. The translational impulse<br />

given to the bat equals the change <strong>in</strong> l<strong>in</strong>ear momentum of the<br />

ball as given by equation 2136 coupledwiththeconservationof<br />

l<strong>in</strong>ear momentum<br />

y<br />

O<br />

C<br />

P = ∆p <br />

= ∆v<br />

<br />

s<br />

M<br />

y<br />

Similarly equation 2142 gives that the angular impulse equals<br />

the change <strong>in</strong> angular momentum about the center of mass to be<br />

0<br />

S<br />

x<br />

T= s × P = ∆L = ∆ω


36 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

The above equations give that<br />

∆v = P <br />

∆ω <br />

= s × P<br />

<br />

Assume that the bat was stationary prior to the strike, then after the strike the net translational velocity<br />

of a po<strong>in</strong>t along the body-fixed symmetry axis of the bat at a distance from the center of mass, is given<br />

by<br />

v () =∆v + ∆ω × y = P + 1 ((s × P) × y) = P + 1 [(s · y) P− (s · P) y]<br />

<br />

It is assumed that and are perpendicular and thus (s · P) =0which simplifies the above equation to<br />

v () =∆v + ∆ω × y = P µ<br />

<br />

(s · y)<br />

1+<br />

<br />

Note that the translational velocity of the location along the bat symmetry axis at a distance from the<br />

center of mass, is zero if the bracket equals zero, that is, if<br />

s · y = − <br />

= −2 <br />

where is called the radius of gyration of the body about the center of mass. Note that when the scalar<br />

product · = − <br />

<br />

= −2 then there will be no translational motion at the po<strong>in</strong>t . This po<strong>in</strong>t on the<br />

axis lies on the opposite side of the center of mass from the strike po<strong>in</strong>t , and is called the center of<br />

percussion correspond<strong>in</strong>g to the impulse at the po<strong>in</strong>t . The center of percussion often is referred to as the<br />

“sweet spot” for an object correspond<strong>in</strong>g to the impulse at the po<strong>in</strong>t . For a baseball bat the batter holds<br />

the bat at the center of percussion so that they do not feel an impulse <strong>in</strong> their hands when the ball is struck<br />

at the po<strong>in</strong>t . This pr<strong>in</strong>ciple is used extensively to design bats for all sports <strong>in</strong>volv<strong>in</strong>g strik<strong>in</strong>g a ball with<br />

a bat, such as, cricket, squash, tennis, etc. as well as weapons such of swords and axes used to decapitate<br />

opponents.<br />

2.15 Example: Energy transfer <strong>in</strong> charged-particle scatter<strong>in</strong>g<br />

Consider a particle of charge + 1 mov<strong>in</strong>g with very high<br />

velocity 0 along a straight l<strong>in</strong>e that passes a distance <br />

from another charge + 2 and mass . F<strong>in</strong>d the energy <br />

transferredtothemass dur<strong>in</strong>g the encounter assum<strong>in</strong>g<br />

the force is given by Coulomb’s law electrostatics. S<strong>in</strong>ce the<br />

charged particle 1 moves at very high speed it is assumed<br />

that charge 2 does not change position dur<strong>in</strong>g the encounter.<br />

Assume that charge 1 moves along the − axis through the<br />

orig<strong>in</strong> while charge 2 is located on the axis at = .<br />

Let us consider the impulse given to charge 2 dur<strong>in</strong>g the<br />

encounter. By symmetry the componentmustcancelwhile<br />

the component is given by<br />

But<br />

= = − 1 2<br />

4 0 2 cos = − 1 2 <br />

cos <br />

4 0 2 <br />

ṗ<br />

O<br />

y<br />

+e 1<br />

V 0<br />

m<br />

Charged-particle scatter<strong>in</strong>g<br />

+e 2<br />

x<br />

˙ = − 0 cos <br />

where<br />

<br />

=cos( − ) =− cos


2.13. SOLUTION OF MANY-BODY EQUATIONS OF MOTION 37<br />

Thus<br />

Integrate from 2 3 2<br />

= − 1 2<br />

4 0 0<br />

cos <br />

gives that the total momentum imparted to 2 is<br />

= − Z 3<br />

1 2 2<br />

4 0 0 <br />

2<br />

Thus the recoil energy of charge 2 is given by<br />

2 = 2 <br />

2 = 1<br />

2<br />

cos = 1 2<br />

2 0 0<br />

µ<br />

1 2<br />

2 0 0<br />

2<br />

2.13 Solution of many-body equations of motion<br />

The follow<strong>in</strong>g are general methods used to solve Newton’s many-body equations of motion for practical<br />

problems.<br />

2.13.1 Analytic solution<br />

In practical problems one has to solve a set of equations of motion s<strong>in</strong>ce the forces depend on the location<br />

of every body <strong>in</strong>volved. For example one may be deal<strong>in</strong>g with a set of coupled oscillators such as the<br />

many components that comprise the suspension system of an automobile. Often the coupled equations of<br />

motion comprise a set of coupled second-order differential equations. The first approach to solve such a<br />

system is to try an analytic solution compris<strong>in</strong>g a general solution of the <strong>in</strong>homogeneous equation plus one<br />

particular solution of the <strong>in</strong>homogeneous equation. Another approach is to employ numeric <strong>in</strong>tegration us<strong>in</strong>g<br />

acomputer.<br />

2.13.2 Successive approximation<br />

When the system of coupled differential equations of motion is too complicated to solve analytically, one<br />

can use the method of successive approximation. The differential equations are transformed to <strong>in</strong>tegral<br />

equations. Then one starts with some <strong>in</strong>itial conditions to make a first order estimate of the functions. The<br />

functions determ<strong>in</strong>ed by this first order estimate then are used <strong>in</strong> a second iteration and this is repeated<br />

until the solution converges. An example of this approach is when mak<strong>in</strong>g Hartree-Foch calculations of the<br />

electron distributions <strong>in</strong> an atom. The first order calculation uses the electron distributions predicted by<br />

the one-electron model of the atom. This result then is used to compute the <strong>in</strong>fluence of the electron charge<br />

distribution around the nucleus on the charge distribution of the atom for a second iteration etc.<br />

2.13.3 Perturbation method<br />

The perturbation technique can be applied if the force separates <strong>in</strong>to two parts = 1 + 2 where 1 2<br />

and the solution is known for the dom<strong>in</strong>ant 1 part of the force. Then the correction to this solution due<br />

to addition of the perturbation 2 usually is easier to evaluate. As an example, consider that one of the<br />

Space Shuttle thrusters fires. In pr<strong>in</strong>ciple one has all the gravitational forces act<strong>in</strong>g plus the thrust force<br />

of the thruster. The perturbation approach is to assume that the trajectory of the Space Shuttle <strong>in</strong> the<br />

earth’s gravitational field is known. Then the perturbation to this motion due to the very small thrust,<br />

produced by the thruster, is evaluated as a small correction to the motion <strong>in</strong> the Earth’s gravitational field.<br />

This perturbation technique is used extensively <strong>in</strong> physics, especially <strong>in</strong> quantum physics. An example<br />

from my own research is scatter<strong>in</strong>g of a 1 208 ion<strong>in</strong>theCoulombfield of a 197 nucleus The<br />

trajectory for elastic scatter<strong>in</strong>g is simple to calculate s<strong>in</strong>ce neither nucleus is excited and the total energy and<br />

momenta are conserved. However, usually one of these nuclei will be <strong>in</strong>ternally excited by the electromagnetic<br />

<strong>in</strong>teraction. This is called Coulomb excitation. The effect of the Coulomb excitation usually can be treated as<br />

a perturbation by assum<strong>in</strong>g that the trajectory is given by the elastic scatter<strong>in</strong>g solution and then calculate<br />

the excitation probability assum<strong>in</strong>g the Coulomb excitation of the nucleus is a small perturbation to the<br />

trajectory.


38 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

2.14 Newton’s Law of Gravitation<br />

Gravitation plays a fundamental role <strong>in</strong> classical mechanics<br />

as well as be<strong>in</strong>g an important example of a conservative<br />

central ¡ ¢<br />

1 2<br />

force. Although you may not be familiar with<br />

use of vector calculus for the gravitational field g, itisassumed<br />

that you have met the identical approach for studies<br />

of the electric field E <strong>in</strong> electrostatics. The primary difference<br />

is that mass replaces charge and gravitational<br />

field g replaces the electric field E. This chapter reviews the<br />

concepts of vector calculus as used for study of conservative<br />

<strong>in</strong>verse-square law central fields.<br />

In 1666 Newton formulated the Theory of Gravitation<br />

which he eventually published <strong>in</strong> the Pr<strong>in</strong>cipia <strong>in</strong> 1687 Newton’s<br />

Law of Gravitation states that each mass particle attracts<br />

every other particle <strong>in</strong> the universe with a force that<br />

varies directly as the product of the mass and <strong>in</strong>versely as<br />

the square of the distance between them. That is, the force<br />

on a gravitational po<strong>in</strong>t mass produced by a mass <br />

F = − <br />

br<br />

2 (2.143)<br />

where br is the unit vector po<strong>in</strong>t<strong>in</strong>g from the gravitational<br />

x<br />

z<br />

r’<br />

dx’dy’dz’<br />

Figure 2.7: Gravitational force on mass m due<br />

to an <strong>in</strong>f<strong>in</strong>itessimal volume element of the mass<br />

density distribution.<br />

mass to the gravitational mass as shown <strong>in</strong> figure 27. Note that the force is attractive, that<br />

is, it po<strong>in</strong>ts toward the other mass. This is <strong>in</strong> contrast to the repulsive electrostatic force between two<br />

similar charges. Newton’s law was verified by Cavendish us<strong>in</strong>g a torsion balance. The experimental value of<br />

=(66726 ± 00008) × 10 −11 · 2 2 <br />

The gravitational force between po<strong>in</strong>t particles can be extended to f<strong>in</strong>ite-sized bodies us<strong>in</strong>g the fact that<br />

the gravitational force field satisfies the superposition pr<strong>in</strong>ciple, that is, the net force is the vector sum of the<br />

<strong>in</strong>dividual forces between the component po<strong>in</strong>t particles. Thus the force summed over the mass distribution<br />

is<br />

F (r) <br />

= − <br />

X<br />

=1<br />

<br />

2 <br />

r<br />

r-r’<br />

br (2.144)<br />

where r is the vector from the gravitational mass to the gravitational mass at the position r.<br />

For a cont<strong>in</strong>uous gravitational mass distribution (r 0 ), the net force on the gravitational mass at<br />

the location r can be written as<br />

Z (r 0 )<br />

³br − br 0´<br />

F (r) =− <br />

(r − r 0 ) 2 0 (2.145)<br />

where 0 is the volume element at the po<strong>in</strong>t r 0 as illustrated <strong>in</strong> figure 27.<br />

2.14.1 Gravitational and <strong>in</strong>ertial mass<br />

Newton’s Laws use the concept of <strong>in</strong>ertial mass ≡ <strong>in</strong> relat<strong>in</strong>g the force F to acceleration a<br />

F = a (2.146)<br />

and momentum p to velocity v<br />

p = v (2.147)<br />

That is, <strong>in</strong>ertial mass is the constant of proportionality relat<strong>in</strong>g the acceleration to the applied force.<br />

The concept of gravitational mass is the constant of proportionality between the gravitational force<br />

and the amount of matter. That is, on the surface of the earth, the gravitational force is assumed to be<br />

F = <br />

"−<br />

X<br />

=1<br />

<br />

2 <br />

br <br />

#<br />

= g (2.148)<br />

m<br />

y


2.14. NEWTON’S LAW OF GRAVITATION 39<br />

where g is the gravitational field which is a position-dependent force per unit gravitational mass po<strong>in</strong>t<strong>in</strong>g<br />

towards the center of the Earth. The gravitational mass is measured when an object is weighed.<br />

Newton’s Law of Gravitation leads to the relation for the gravitational field g (r) at the location r due<br />

to a gravitational mass distribution at the location r 0 as given by the <strong>in</strong>tegral over the gravitational mass<br />

density <br />

Z<br />

g (r) =−<br />

<br />

(r 0 )<br />

³br − br 0´<br />

(r − r 0 ) 2 0 (2.149)<br />

The acceleration of matter <strong>in</strong> a gravitational field relates the gravitational and <strong>in</strong>ertial masses<br />

F = g = a (2.150)<br />

Thus<br />

a = <br />

g (2.151)<br />

<br />

That is, the acceleration of a body depends on the gravitational strength and the ratio of the gravitational<br />

and <strong>in</strong>ertial masses. It has been shown experimentally that all matter is subject to the same acceleration<br />

<strong>in</strong> vacuum at a given location <strong>in</strong> a gravitational field. That is, <br />

<br />

is a constant common to all materials.<br />

Galileo first showed this when he dropped objects from the Tower of Pisa. Modern experiments have shown<br />

that this is true to 5 parts <strong>in</strong> 10 13 .<br />

The exact equivalence of gravitational mass and <strong>in</strong>ertial mass is called the weak pr<strong>in</strong>ciple of equivalence<br />

which underlies the General Theory of Relativity as discussed <strong>in</strong> chapter 17. Itisconvenienttouse<br />

the same unit for the gravitational and <strong>in</strong>ertial masses and thus they both can be written <strong>in</strong> terms of the<br />

common mass symbol .<br />

= = (2.152)<br />

Therefore the subscripts and can be omitted <strong>in</strong> equations 2150 and 2152. Also the local acceleration<br />

due to gravity a can be written as<br />

a = g (2.153)<br />

The gravitational field g ≡ F has units of <strong>in</strong> the MKS system while the acceleration a has units 2 .<br />

2.14.2 Gravitational potential energy <br />

Chapter 2102 showed that a conservative field can be expressed <strong>in</strong><br />

termsoftheconceptofapotentialenergy(r) which depends on<br />

position. The potential energy difference ∆ → between two po<strong>in</strong>ts<br />

r and r , is the work done mov<strong>in</strong>g from to aga<strong>in</strong>st a force F. That<br />

is:<br />

Z <br />

∆ → = (r ) − (r )=− F · l (2.154)<br />

<br />

In general, this l<strong>in</strong>e <strong>in</strong>tegral depends on the path taken.<br />

Consider the gravitational field produced by a s<strong>in</strong>gle po<strong>in</strong>t mass<br />

1 The work done mov<strong>in</strong>g a mass 0 from to <strong>in</strong> this gravitational<br />

field can be calculated along an arbitrary path shown <strong>in</strong> figure<br />

28 by assum<strong>in</strong>g Newton’s law of gravitation. Then the force on 0<br />

due to po<strong>in</strong>t mass 1 is;<br />

mg<br />

F<br />

dl<br />

F = − 1 0<br />

br<br />

2 (2.155)<br />

Express<strong>in</strong>g l <strong>in</strong> spherical coord<strong>in</strong>ates l =ˆr+ˆθ+ s<strong>in</strong> ˆφ gives<br />

the path <strong>in</strong>tegral (2154) from ( ) to ( ) is<br />

Z <br />

Z <br />

h<br />

∆ → = − F · l =<br />

<br />

<br />

∙ 1<br />

= − 1 0 − 1 ¸<br />

<br />

1 0<br />

2<br />

(ˆr·br + ˆr · ˆθ + s<strong>in</strong> ˆr · ˆφ)<br />

i<br />

= <br />

Figure 2.8: Work done aga<strong>in</strong>st a<br />

force field mov<strong>in</strong>g from a to b.<br />

Z <br />

<br />

1 0<br />

br<br />

2 · br<br />

(2.156)


40 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

s<strong>in</strong>ce the scalar product of the unit vectors br · br =1 Note that the second two terms also cancel s<strong>in</strong>ce<br />

br · ˆθ = ˆr · ˆφ =0s<strong>in</strong>ce the unit vectors are mutually orthogonal. Thusthel<strong>in</strong>e<strong>in</strong>tegraljustdependsonlyon<br />

the start<strong>in</strong>g and end<strong>in</strong>g radii and is <strong>in</strong>dependent of the angular coord<strong>in</strong>ates or the detailed path taken between<br />

( ) and ( ) <br />

Consider the Pr<strong>in</strong>ciple of Superposition for a gravitational field produced by a set of po<strong>in</strong>t masses. The<br />

l<strong>in</strong>e <strong>in</strong>tegral then can be written as:<br />

Z <br />

X<br />

Z <br />

X<br />

∆→ = − F · l = − F · l = ∆→ (2.157)<br />

<br />

=1 =1<br />

Thus the net potential energy difference is the sum of the contributions from each po<strong>in</strong>t mass produc<strong>in</strong>g the<br />

gravitational force field. S<strong>in</strong>ce each component is conservative, then the total potential energy difference also<br />

must be conservative. For a conservative force, this l<strong>in</strong>e <strong>in</strong>tegral is <strong>in</strong>dependent of the path taken, itdepends<br />

only on the start<strong>in</strong>g and end<strong>in</strong>g positions, r and r . That is, the potential energy is a local function<br />

dependent only on position. The usefulness of gravitational potential energy is that, s<strong>in</strong>ce the gravitational<br />

force is a conservative force, it is possible to solve many problems <strong>in</strong> classical mechanics us<strong>in</strong>g the fact<br />

that the sum of the k<strong>in</strong>etic energy and potential energy is a constant. Note that the gravitational field is<br />

conservative, s<strong>in</strong>ce the potential energy difference ∆→ is <strong>in</strong>dependent of the path taken. It is conservative<br />

becausetheforceisradial and time <strong>in</strong>dependent, it is not due to the 1 <br />

dependence of the field.<br />

2<br />

2.14.3 Gravitational potential <br />

Us<strong>in</strong>g F = 0 g gives that the change <strong>in</strong> potential energy due to mov<strong>in</strong>g a mass 0 from to <strong>in</strong> a<br />

gravitational field g is:<br />

∆ <br />

→ = − 0<br />

Z <br />

<br />

g · l (2.158)<br />

Note that the probe mass 0 factors out from the <strong>in</strong>tegral. It is convenient to def<strong>in</strong>e a new quantity called<br />

gravitational potential where<br />

∆ <br />

→ =<br />

<br />

Z<br />

∆<br />

<br />

→<br />

= − g · l (2.159)<br />

0<br />

<br />

That is; gravitational potential differenceistheworkthatmustbedone,perunitmass,tomovefroma to<br />

b with no change <strong>in</strong> k<strong>in</strong>etic energy. Be careful not to confuse the gravitational potential energy difference<br />

∆ → and gravitational potential difference ∆ → ,thatis,∆ has units of energy, ,while∆ has<br />

units of .<br />

The gravitational potential is a property of the gravitational force field; it is given as m<strong>in</strong>us the l<strong>in</strong>e<br />

<strong>in</strong>tegral of the gravitational field from to . The change <strong>in</strong> gravitational potential energy for mov<strong>in</strong>g a<br />

mass 0 from to is given <strong>in</strong> terms of gravitational potential by:<br />

Superposition and potential<br />

∆ <br />

→ = 0 ∆ <br />

→ (2.160)<br />

Previously it was shown that the gravitational force is conservative for the superposition of many masses.<br />

To recap, if the gravitational field<br />

g = g 1 + g 2 + g 3 (2.161)<br />

then<br />

Z <br />

Z <br />

Z <br />

Z <br />

<br />

→ = − g · l = − g 1 · l − g 2 · l − g 3 · l = Σ → (2.162)<br />

<br />

<br />

<br />

<br />

Thus gravitational potential is a simple additive scalar field because the Pr<strong>in</strong>ciple of Superposition applies.<br />

The gravitational potential, between two po<strong>in</strong>ts differ<strong>in</strong>g by <strong>in</strong> height, is . Clearly, the greater or ,<br />

the greater the energy released by the gravitational field when dropp<strong>in</strong>g a body through the height . The<br />

unit of gravitational potential is the


2.14. NEWTON’S LAW OF GRAVITATION 41<br />

2.14.4 Potential theory<br />

The gravitational force and electrostatic force both obey the <strong>in</strong>verse square law, for which the field and<br />

correspond<strong>in</strong>g potential are related by:<br />

Z <br />

∆ → = − g · l (2.163)<br />

<br />

For an arbitrary <strong>in</strong>f<strong>in</strong>itessimal element distance l the change <strong>in</strong> gravitational potential is<br />

= −g · l (2.164)<br />

Us<strong>in</strong>g cartesian coord<strong>in</strong>ates both g and l can be written as<br />

g = b i + b j + k b l = b i + b j + k b (2.165)<br />

Tak<strong>in</strong>g the scalar product gives:<br />

= −g · l = − − − (2.166)<br />

Differential calculus expresses the change <strong>in</strong> potential <strong>in</strong> terms of partial derivatives by:<br />

By association, 2166 and 2167 imply that<br />

= <br />

<br />

+<br />

<br />

<br />

<br />

+ (2.167)<br />

<br />

= − <br />

<br />

= − <br />

<br />

= − <br />

<br />

(2.168)<br />

Thus on each axis, the gravitational field can be written as m<strong>in</strong>us the gradient of the gravitational potential.<br />

In three dimensions, the gravitational field is m<strong>in</strong>us the total gradient of potential and the gradient of the<br />

scalar function canbewrittenas:<br />

g = −∇ (2.169)<br />

In cartesian coord<strong>in</strong>ates this equals<br />

∙<br />

g = − b<br />

i<br />

+b j <br />

+ k b ¸<br />

(2.170)<br />

<br />

Thus the gravitational field is just the gradient of the gravitational potential, which always is perpendicular<br />

to the equipotentials. Skiers are familiar with the concept of gravitational equipotentials and the fact that<br />

the l<strong>in</strong>e of steepest descent, and thus maximum acceleration, is perpendicular to gravitational equipotentials<br />

of constant height. The advantage of us<strong>in</strong>g potential theory for <strong>in</strong>verse-square law forces is that scalar<br />

potentials replace the more complicated vector forces, which greatly simplifies calculation. Potential theory<br />

plays a crucial role for handl<strong>in</strong>g both gravitational and electrostatic forces.<br />

2.14.5 Curl of the gravitational field<br />

It has been shown that the gravitational field is conservative, that is<br />

∆ → is <strong>in</strong>dependent of the path taken between and . Therefore,<br />

equation 2159 gives that the gravitational potential is <strong>in</strong>dependent of<br />

the path taken between two po<strong>in</strong>ts and . Consider two possible paths<br />

between and as shown <strong>in</strong> figure 29. The l<strong>in</strong>e <strong>in</strong>tegral from to via<br />

route 1 is equal and opposite to the l<strong>in</strong>e <strong>in</strong>tegral back from to via<br />

route 2 if the gravitational field is conservative as shown earlier.<br />

A better way of express<strong>in</strong>g this is that the l<strong>in</strong>e <strong>in</strong>tegral of the gravitational<br />

field is zero around any closed path. Thus the l<strong>in</strong>e <strong>in</strong>tegral between<br />

and , viapath1, and return<strong>in</strong>g back to , viapath2, areequaland<br />

opposite. That is, the net l<strong>in</strong>e <strong>in</strong>tegral for a closed loop is zero<br />

1<br />

2<br />

Figure 2.9: Circulation of the<br />

gravitational field.


42 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

I<br />

g · l =0 (2.171)<br />

which is a measure of the circulation of the gravitational field. The fact that the circulation equals zero<br />

corresponds to the statement that the gravitational field is radial for a po<strong>in</strong>t mass.<br />

Stokes Theorem, discussed <strong>in</strong> appendix 3 states that<br />

I Z<br />

<br />

F · l =<br />

<br />

<br />

<br />

<br />

(∇ × F) · S (2.172)<br />

Thus the zero circulation of the gravitational field can be rewritten as<br />

I Z<br />

g · l = (∇ × g) · S =0 (2.173)<br />

<br />

<br />

<br />

<br />

<br />

S<strong>in</strong>ce this is <strong>in</strong>dependent of the shape of the perimeter , therefore<br />

∇ × g =0 (2.174)<br />

That is, the gravitational field is a curl-free field.<br />

A property of any curl-free field is that it can be expressed as the gradient of a scalar potential s<strong>in</strong>ce<br />

∇ × ∇ =0 (2.175)<br />

Therefore, the curl-free gravitational field can be related to a scalar potential as<br />

g = −∇ (2.176)<br />

Thus is consistent with the above def<strong>in</strong>ition of gravitational potential <strong>in</strong> that the scalar product<br />

Z Z <br />

Z X<br />

Z<br />

<br />

<br />

∆ → = − g · l = (∇) · l =<br />

= (2.177)<br />

<br />

<br />

<br />

An identical relation between the electric field and electric potential applies for the <strong>in</strong>verse-square law<br />

electrostatic field.<br />

Reference potentials:<br />

Note that only differences <strong>in</strong> potential energy, , and gravitational potential, , are mean<strong>in</strong>gful, the absolute<br />

values depend on some arbitrarily chosen reference. However, often it is useful to measure gravitational<br />

potential with respect to a particular arbitrarily chosen reference po<strong>in</strong>t such as to sea level. Aircraft<br />

pilots are required to set their altimeters to read with respect to sea level rather than their departure<br />

airport. This ensures that aircraft leav<strong>in</strong>g from say both Rochester, 559 0 and Denver 5000 0 , have<br />

their altimeters set to a common reference to ensure that they do not collide. The gravitational force is the<br />

gradient of the gravitational field which only depends on differences <strong>in</strong> potential, and thus is <strong>in</strong>dependent of<br />

any constant reference.<br />

Gravitational potential due to cont<strong>in</strong>uous distributions of charge Suppose mass is distributed<br />

over a volume with a density at any po<strong>in</strong>t with<strong>in</strong> the volume. The gravitational potential at any field<br />

po<strong>in</strong>t due to an element of mass = at the po<strong>in</strong>t 0 is given by:<br />

Z<br />

( 0 ) 0<br />

∆ ∞→ = −<br />

(2.178)<br />

0 <br />

This <strong>in</strong>tegral is over a scalar quantity. S<strong>in</strong>ce gravitational potential is a scalar quantity, it is easier to<br />

compute than is the vector gravitational field g . If the scalar potential field is known, then the gravitational<br />

field is derived by tak<strong>in</strong>g the gradient of the gravitational potential.


2.14. NEWTON’S LAW OF GRAVITATION 43<br />

2.14.6 Gauss’s Law for Gravitation<br />

The flux Φ of the gravitational field g through a surface<br />

, asshown<strong>in</strong>figure 210, isdef<strong>in</strong>ed as<br />

Z<br />

Φ ≡ g · S (2.179)<br />

<br />

Note that there are two possible perpendicular directions<br />

that could be chosen for the surface vector S Us<strong>in</strong>g<br />

g<br />

Newton’s law of gravitation for a po<strong>in</strong>t mass the flux<br />

through the surface is<br />

Z<br />

br · S<br />

Φ = −<br />

2 (2.180)<br />

Note that the solid angle subtended by the surface <br />

at an angle to the normal from the po<strong>in</strong>t mass is given<br />

by<br />

cos br · S<br />

Ω =<br />

2 =<br />

2 (2.181)<br />

Thus the net gravitational flux equals<br />

Z<br />

Figure 2.10: Flux of the gravitational field through<br />

an <strong>in</strong>f<strong>in</strong>itessimal surface element dS.<br />

Φ = − Ω (2.182)<br />

<br />

Consider a closed surface where the direction of the surface vector S is def<strong>in</strong>ed as outwards. The net<br />

flux out of this closed surface is given by<br />

I<br />

I<br />

br · S<br />

Φ = −<br />

2 = − Ω = −4 (2.183)<br />

<br />

This is <strong>in</strong>dependent of where the po<strong>in</strong>t mass lies with<strong>in</strong> the closed surface or on the shape of the closed<br />

surface. Note that the solid angle subtended is zero if the po<strong>in</strong>t mass lies outside the closed surface. Thus<br />

the flux is as given by equation 2183 if the mass is enclosed by the closed surface, while it is zero if the mass<br />

is outside of the closed surface.<br />

S<strong>in</strong>ce the flux for a po<strong>in</strong>t mass is <strong>in</strong>dependent of the location of the mass with<strong>in</strong> the volume enclosed by<br />

the closed surface, and us<strong>in</strong>g the pr<strong>in</strong>ciple of superposition for the gravitational field, then for enclosed<br />

po<strong>in</strong>t masses the net flux is<br />

Z<br />

X<br />

Φ ≡ g · S = −4 (2.184)<br />

<br />

This can be extended to cont<strong>in</strong>uous mass distributions, with local mass density giv<strong>in</strong>g that the net flux<br />

Z<br />

Z<br />

Φ ≡ g · S = −4 (2.185)<br />

<br />

<br />

<br />

<br />

Gauss’s Divergence Theorem was given <strong>in</strong> appendix 2 as<br />

I Z<br />

Φ = F · S = ∇ · F (2.186)<br />

<br />

<br />

<br />

Apply<strong>in</strong>g the Divergence Theorem to Gauss’s law gives that<br />

I Z<br />

Z<br />

Φ = g · S = ∇ · g = −4<br />

or<br />

<br />

Z<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

[∇ · g +4] =0 (2.187)<br />

dS


44 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

This is true <strong>in</strong>dependent of the shape of the surface, thus the divergence of the gravitational field<br />

∇ · g = −4 (2.188)<br />

This is a statement that the gravitational field of a po<strong>in</strong>t mass has a 1<br />

<br />

dependence.<br />

2<br />

Us<strong>in</strong>g the fact that the gravitational field is conservative, this can be expressed as the gradient of the<br />

gravitational potential <br />

g = −∇ (2.189)<br />

and Gauss’s law, then becomes<br />

which also can be written as Poisson’s equation<br />

∇ · ∇ =4 (2.190)<br />

∇ 2 =4 (2.191)<br />

Know<strong>in</strong>g the mass distribution allows determ<strong>in</strong>ation of the potential by solv<strong>in</strong>g Poisson’s equation.<br />

A special case that often is encountered is when the mass distribution is zero <strong>in</strong> a given region. Then the<br />

potential for this region can be determ<strong>in</strong>ed by solv<strong>in</strong>g Laplace’s equation with known boundary conditions.<br />

∇ 2 =0 (2.192)<br />

For example, Laplace’s equation applies <strong>in</strong> the free space between the masses. It is used extensively <strong>in</strong> electrostatics<br />

to compute the electric potential between charged conductors which themselves are equipotentials.<br />

2.14.7 Condensed forms of Newton’s Law of Gravitation<br />

The above discussion has resulted <strong>in</strong> several alternative expressions of Newton’s Law of Gravitation that will<br />

be summarized here. The most direct statement of Newton’s law is<br />

Z<br />

g (r) =−<br />

<br />

(r 0 )<br />

³br − br 0´<br />

(r − r 0 ) 2 0 (2.193)<br />

An elegant way to express Newton’s Law of Gravitation is <strong>in</strong> terms of the flux and circulation of the<br />

gravitational field. That is,<br />

Flux:<br />

Z<br />

Circulation:<br />

Φ ≡<br />

<br />

Z<br />

g · S = −4<br />

I<br />

<br />

<br />

(2.194)<br />

g · l =0 (2.195)<br />

The flux and circulation are better expressed <strong>in</strong> terms of the vector differential concepts of divergence<br />

and curl.<br />

Divergence:<br />

∇ · g = −4 (2.196)<br />

Curl:<br />

∇ × g =0 (2.197)<br />

Remember that the flux and divergence of the gravitational field are statements that the field between<br />

po<strong>in</strong>t masses has a 1<br />

<br />

dependence. The circulation and curl are statements that the field between po<strong>in</strong>t<br />

2<br />

masses is radial.<br />

Because the gravitational field is conservative it is possible to use the concept of the scalar potential<br />

field This concept is especially useful for solv<strong>in</strong>g some problems s<strong>in</strong>ce the gravitational potential can be<br />

evaluated us<strong>in</strong>g the scalar <strong>in</strong>tegral<br />

Z<br />

( 0 ) 0<br />

∆ ∞→ = −<br />

(2.198)<br />

0


2.14. NEWTON’S LAW OF GRAVITATION 45<br />

An alternate approach is to solve Poisson’s equation if the boundary values and mass distributions are known<br />

where Poisson’s equation is:<br />

∇ 2 =4 (2.199)<br />

These alternate expressions of Newton’s law of gravitation can be exploited to solve problems.<br />

method of solution is identical to that used <strong>in</strong> electrostatics.<br />

The<br />

2.16 Example: Field of a uniform sphere<br />

Consider the simple case of the gravitational field due to a uniform sphere of matter of radius and<br />

mass . Then the volume mass density<br />

=<br />

3<br />

4 3<br />

The gravitational field and potential for this uniform sphere of matter can be derived three ways;<br />

a) The field can be evaluated by directly <strong>in</strong>tegrat<strong>in</strong>g over the volume<br />

Z (r 0 )<br />

³br − br 0´<br />

g (r) =−<br />

(r − r 0 ) 2 0<br />

b) The potential can be evaluated directly by <strong>in</strong>tegration of<br />

Z<br />

( 0 ) 0<br />

∆ ∞→ = −<br />

0 <br />

and then<br />

g = −∇<br />

c) The obvious spherical symmetry can be used <strong>in</strong> conjunction<br />

with Gauss’s law to easily solve this problem.<br />

Z<br />

Z<br />

g · S = −4 <br />

<br />

<br />

<br />

g<br />

0<br />

-GM r<br />

-GM<br />

r²<br />

4 2 () =−4<br />

(rR)<br />

That is: for <br />

g = − 2 br<br />

(rR)<br />

-GM | 3R²-r² |<br />

-GM<br />

r<br />

Similarly, for <br />

4 2 () = 4 3 3 <br />

(rR)<br />

Gravitational field g and gravitational<br />

potential Φ of a uniformly-dense<br />

spherical mass distribution of radius .<br />

That is:<br />

g = − 3 r<br />

(rR)<br />

The field <strong>in</strong>side the Earth is radial and is proportional to the distance from the center of the Earth. This<br />

is Hooke’s Law, and thus ignor<strong>in</strong>g air drag, any body dropped down a qhole through the center of the Earth<br />

<br />

will undergo harmonic oscillations with an angular frequency of 0 =<br />

<br />

= p <br />

3 This gives a period of<br />

oscillation of 14 hours, which is about the length of a 235 lecture <strong>in</strong> classical mechanics, which may seem<br />

like a long time.<br />

Clearly method (c) is much simpler to solve for this case. In general, look for a symmetry that allows<br />

identification of a surface upon which the magnitude and direction of the fieldisconstant. Forsuchcases<br />

use Gauss’s law. Otherwise use methods (a) or (b) whichever one is easiest to apply. Further examples will<br />

not be given here s<strong>in</strong>ce they are essentially identical to those discussed extensively <strong>in</strong> electrostatics.


46 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

2.15 Summary<br />

Newton’s Laws of Motion:<br />

A cursory review of Newtonian mechanics has been presented. The concept of <strong>in</strong>ertial frames of reference<br />

was <strong>in</strong>troduced s<strong>in</strong>ce Newton’s laws of motion apply only to <strong>in</strong>ertial frames of reference.<br />

Newton’s Law of motion<br />

F = p<br />

(26)<br />

<br />

leads to second-order equations of motion which can be difficult to handle for many-body systems.<br />

Solution of Newton’s second-order equations of motion can be simplified us<strong>in</strong>g the three first-order <strong>in</strong>tegrals<br />

coupled with correspond<strong>in</strong>g conservation laws. The first-order time <strong>in</strong>tegral for l<strong>in</strong>ear momentum<br />

is Z 2<br />

F =<br />

1<br />

Z 2<br />

The first-order time <strong>in</strong>tegral for angular momentum is<br />

L <br />

<br />

= r × p <br />

= N <br />

1<br />

p <br />

=(p 2 − p 1 ) <br />

Z 2<br />

1<br />

N =<br />

Z 2<br />

1<br />

L <br />

=(L 2 − L 1 ) <br />

The first-order spatial <strong>in</strong>tegral is related to k<strong>in</strong>etic energy and the concept of work. That is<br />

(210)<br />

(216)<br />

F = Z 2<br />

<br />

F · r =( 2 − 1 ) <br />

r <br />

<br />

1<br />

(221)<br />

The conditions that lead to conservation of l<strong>in</strong>ear and angular momentum and total mechanical energy<br />

were discussed for many-body systems. The important class of conservative forces was shown to apply if<br />

the position-dependent force do not depend on time or velocity, and if the work done by a force R 2<br />

1 F · r <br />

is <strong>in</strong>dependent of the path taken between the <strong>in</strong>itial and f<strong>in</strong>al locations. The total mechanical energy is a<br />

constant of motion when the forces are conservative.<br />

It was shown that the concept of center of mass of a many-body or f<strong>in</strong>ite sized body separates naturally<br />

for all three first-order <strong>in</strong>tegrals. The center of mass is that po<strong>in</strong>t about which<br />

X<br />

Z<br />

r 0 = r 0 =0<br />

(Centre of mass def<strong>in</strong>ition)<br />

where r 0 is the vector def<strong>in</strong><strong>in</strong>g the location of mass with respect to the center of mass. The concept of<br />

center of mass greatly simplifies the description of the motion of f<strong>in</strong>ite-sized bodies and many-body systems<br />

by separat<strong>in</strong>g out the important <strong>in</strong>ternal <strong>in</strong>teractions and correspond<strong>in</strong>g underly<strong>in</strong>g physics, from the trivial<br />

overall translational motion of a many-body system..<br />

The Virial theorem states that the time-averaged properties are related by<br />

* +<br />

h i = − 1 X<br />

F · r (286)<br />

2<br />

<br />

It was shown that the Virial theorem is useful for relat<strong>in</strong>g the time-averaged k<strong>in</strong>etic and potential energies,<br />

especially for cases <strong>in</strong>volv<strong>in</strong>g either l<strong>in</strong>ear or <strong>in</strong>verse-square forces.<br />

Typical examples were presented of application of Newton’s equations of motion to solv<strong>in</strong>g systems<br />

<strong>in</strong>volv<strong>in</strong>g constant, l<strong>in</strong>ear, position-dependent, velocity-dependent, and time-dependent forces, to constra<strong>in</strong>ed<br />

and unconstra<strong>in</strong>ed systems, as well as systems with variable mass. Rigid-body rotation about a body-fixed<br />

rotation axis also was discussed.<br />

It is important to be cognizant of the follow<strong>in</strong>g limitations that apply to Newton’s laws of motion:<br />

1) Newtonian mechanics assumes that all observables are measured to unlimited precision, that is <br />

p r are known exactly. Quantum physics <strong>in</strong>troduces limits to measurement due to wave-particle duality.<br />

2) The Newtonian view is that time and position are absolute concepts. The Theory of Relativity shows<br />

that this is not true. Fortunately for most problems and thus Newtonian mechanics is an excellent<br />

approximation.


2.15. SUMMARY 47<br />

3) Another limitation, to be discussed later, is that it is impractical to solve the equations of motion for<br />

many <strong>in</strong>teract<strong>in</strong>g bodies such as all the molecules <strong>in</strong> a gas. Then it is necessary to resort to us<strong>in</strong>g statistical<br />

averages, this approach is called statistical mechanics.<br />

Newton’s work constitutes a theory of motion <strong>in</strong> the universe that <strong>in</strong>troduces the concept of causality.<br />

Causality is that there is a one-to-one correspondence between cause of effect. Each force causes a known<br />

effect that can be calculated. Thus the causal universe is pictured by philosophers to be a giant mach<strong>in</strong>e<br />

whose parts move like clockwork <strong>in</strong> a predictable and predeterm<strong>in</strong>ed way accord<strong>in</strong>g to the laws of nature. This<br />

is a determ<strong>in</strong>istic view of nature. There are philosophical problems <strong>in</strong> that such a determ<strong>in</strong>istic viewpo<strong>in</strong>t<br />

appears to be contrary to free will. That is, taken to the extreme it implies that you were predest<strong>in</strong>ed to<br />

read this book because it is a natural consequence of this mechanical universe!<br />

Newton’s Laws of Gravitation<br />

Newton’s Laws of Gravitation and the Laws of Electrostatics are essentially identical s<strong>in</strong>ce they both<br />

<strong>in</strong>volve a central <strong>in</strong>verse square-law dependence of the forces. The important difference is that the gravitational<br />

force is attractive whereas the electrostatic force between identical charges is repulsive. That is,<br />

the gravitational constant is replaced by − 1<br />

4 0<br />

,andthemassdensity becomes the charge density for<br />

the case of electrostatics. As a consequence it is unnecessary to make a detailed study of Newton’s law of<br />

gravitation s<strong>in</strong>ce it is identical to what has already been studied <strong>in</strong> your accompany<strong>in</strong>g electrostatic courses.<br />

Table 21 summarizes and compares the laws of gravitation and electrostatics. For both gravitation and<br />

electrostatics the field is central and conservative and depends as 1<br />

2ˆr<br />

The laws of gravitation and electrostatics can be expressed <strong>in</strong> a more useful form <strong>in</strong> terms of the flux and<br />

circulation of the gravitational field as given either <strong>in</strong> the vector <strong>in</strong>tegral or vector differential forms. The<br />

radial <strong>in</strong>dependence of the flux, and correspond<strong>in</strong>g divergence, is a statement that the fields are radial and<br />

have a 1<br />

2ˆr dependence. The statement that the circulation, and correspond<strong>in</strong>g curl, are zero is a statement<br />

that the fields are radial and conservative.<br />

Table 21; Comparison of Newton’s law of gravitation and electrostatics.<br />

Gravitation<br />

Electrostatics<br />

Force field g ≡ F <br />

E ≡ F<br />

<br />

Density Mass density (r 0 ) Charge density (r 0 )<br />

Conservative central field g (r) =− R (r 0 )(r− r 0 )<br />

0 E (r) = 1 (r<br />

(r−r 0 ) 2 4 0<br />

R<br />

)(r− r 0 )<br />

0<br />

(r−r 0 ) 2<br />

Flux<br />

Φ ≡ R g · S = −4 R Φ ≡ R E · S = 1 <br />

<br />

0<br />

R <br />

I<br />

I<br />

<br />

Circulation<br />

g · l =0<br />

E · l =0<br />

Divergence ∇ · g = −4 ∇ · E = 1 0<br />

<br />

Curl ∇ × g =0 ∇ × E =0<br />

Potential<br />

∆ ∞→ = − R ( 0 ) 0<br />

<br />

∆<br />

0 ∞→ = 1<br />

4 0<br />

R<br />

Poisson’s equation ∇ 2 =4 ∇ 2 = − 1 0<br />

<br />

( 0 ) 0<br />

0 <br />

Both the gravitational and electrostatic central fields are conservative mak<strong>in</strong>g it possible to use the<br />

concept of the scalar potential field This concept is especially useful for solv<strong>in</strong>g some problems s<strong>in</strong>ce the<br />

potential can be evaluated us<strong>in</strong>g a scalar <strong>in</strong>tegral. An alternate approach is to solve Poisson’s equation if the<br />

boundary values and mass distributions are known. The methods of solution of Newton’s law of gravitation<br />

are identical to those used <strong>in</strong> electrostatics and are readily accessible <strong>in</strong> the literature.


48 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

Problems<br />

1. Two particles are projected from the same po<strong>in</strong>t with velocities 1 and 2 ,atelevations 1 and 2 , respectively<br />

( 1 2 ). Show that if they are to collide <strong>in</strong> mid-air the <strong>in</strong>terval between the fir<strong>in</strong>gs must be<br />

2 1 2 s<strong>in</strong>( 1 − 2 )<br />

( 1 cos 1 + 2 cos 2 ) <br />

2. The teeter totter comprises two identical weights which hang on droop<strong>in</strong>g arms attached to a peg as shown.<br />

The arrangement is unexpectedly stable and can be spun and rocked with little danger of toppl<strong>in</strong>g over.<br />

m<br />

l<br />

L<br />

l<br />

m<br />

(a) F<strong>in</strong>d an expression for the potential energy of the teeter toy as a function of when the teeter toy is<br />

cocked at an angle about the pivot po<strong>in</strong>t. For simplicity, consider only rock<strong>in</strong>g motion <strong>in</strong> the vertical<br />

plane.<br />

(b) Determ<strong>in</strong>e the equilibrium values(s) of .<br />

(c) Determ<strong>in</strong>e whether the equilibrium is stable, unstable, or neutral for the value(s) of found <strong>in</strong> part (b).<br />

(d) How could you determ<strong>in</strong>e the answers to parts (b) and (c) from a graph of the potential energy versus ?<br />

(e) Expand the expression for the potential energy about = 0 and determ<strong>in</strong>e the frequency of small<br />

oscillations.<br />

3. A particle of mass is constra<strong>in</strong>ed to move on the frictionless <strong>in</strong>ner surface of a cone of half-angle .<br />

(a) F<strong>in</strong>d the restrictions on the <strong>in</strong>itial conditions such that the particle moves <strong>in</strong> a circular orbit about the<br />

vertical axis.<br />

(b) Determ<strong>in</strong>e whether this k<strong>in</strong>d of orbit is stable.<br />

4. Consider a th<strong>in</strong> rod of length and mass .<br />

(a) Draw gravitational field l<strong>in</strong>es and equipotential l<strong>in</strong>es for the rod. What can you say about the equipotential<br />

surfaces of the rod?<br />

(b) Calculate the gravitational potential at a po<strong>in</strong>t that is a distance from one end of the rod and <strong>in</strong> a<br />

direction perpendicular to the rod.<br />

(c) Calculate the gravitational field at by direct <strong>in</strong>tegration.<br />

(d) Could you have used Gauss’s law to f<strong>in</strong>d the gravitational field at ?Whyorwhynot?<br />

5. Consider a s<strong>in</strong>gle particle of mass <br />

(a) Determ<strong>in</strong>e the position and velocity of a particle <strong>in</strong> spherical coord<strong>in</strong>ates.<br />

(b) Determ<strong>in</strong>e the total mechanical energy of the particle <strong>in</strong> potential .<br />

(c) Assume the force is conservative. Show that = −∇ .ShowthatitagreeswithStoke’stheorem.<br />

(d) Show that the angular momentum = × of the particle is conserved.<br />

× dB<br />

d + dA<br />

d × .<br />

H<strong>in</strong>t:<br />

d<br />

d<br />

( × ) =


2.15. SUMMARY 49<br />

6. Consider a fluidwithdensity and velocity <strong>in</strong> some volume . The mass current = determ<strong>in</strong>es the<br />

amount of mass exit<strong>in</strong>g the surface per unit time by the <strong>in</strong>tegral · <br />

(a) Us<strong>in</strong>g the divergence theorem, prove the cont<strong>in</strong>uity equation, ∇ · + <br />

=0<br />

7. A rocket of <strong>in</strong>itial mass burns fuel at constant rate (kilograms per second), produc<strong>in</strong>g a constant force .<br />

The total mass of available fuel is . Assume the rocket starts from rest and moves <strong>in</strong> a fixed direction with<br />

no external forces act<strong>in</strong>g on it.<br />

(a) Determ<strong>in</strong>e the equation of motion of the rocket.<br />

(b) Determ<strong>in</strong>e the f<strong>in</strong>al velocity of the rocket.<br />

(c) Determ<strong>in</strong>e the displacement of the rocket <strong>in</strong> time.<br />

8. Consider a solid hemisphere of radius . Compute the coord<strong>in</strong>ates of the center of mass relative to the center<br />

of the spherical surface used to def<strong>in</strong>e the hemisphere.<br />

9. A 2000kg Ford was travell<strong>in</strong>g south on Mt. Hope Avenue when it collided with your 1000kg sports car travell<strong>in</strong>g<br />

west on Elmwood Avenue. The two badly-damaged cars became entangled <strong>in</strong> the collision and leave a skid mark<br />

that is 20 meters long <strong>in</strong> a direction 14 ◦ to the west of the orig<strong>in</strong>al direction of travel of the Excursion. The<br />

wealthy Excursion driver hires a high-powered lawyer who accuses you of speed<strong>in</strong>g through the <strong>in</strong>tersection.<br />

Use your P235 knowledge, plus the police officer’s report of the recoil direction, the skid length, and knowledge<br />

that the coefficient of slid<strong>in</strong>g friction between the tires and road is =06, to deduce the orig<strong>in</strong>al velocities of<br />

both cars. Were either of the cars exceed<strong>in</strong>g the 30mph speed limit?<br />

10. A particle of mass mov<strong>in</strong>g <strong>in</strong> one dimension has potential energy () = 0 [2( )2 − ( )4 ] where 0 and <br />

are positive constants.<br />

a) F<strong>in</strong>d the force () that acts on the particle.<br />

b) Sketch (). F<strong>in</strong>d the positions of stable and unstable equilibrium.<br />

c) What is the angular frequency of oscillations about the po<strong>in</strong>t of stable equilibrium?<br />

d) What is the m<strong>in</strong>imum speed the particle must have at the orig<strong>in</strong> to escape to <strong>in</strong>f<strong>in</strong>ity?<br />

e) At =0the particle is at the orig<strong>in</strong> and its velocity is positive and equal to the escape velocity. F<strong>in</strong>d ()<br />

andsketchtheresult.<br />

11. a) Consider a s<strong>in</strong>gle-stage rocket travell<strong>in</strong>g <strong>in</strong> a straight l<strong>in</strong>e subject to an external force act<strong>in</strong>g along the<br />

same l<strong>in</strong>e where is the exhaust velocity of the ejected fuel relative to the rocket. Show that the equation of<br />

motion is<br />

˙ = − ˙ + <br />

b) Specialize to the case of a rocket tak<strong>in</strong>g off vertically from rest <strong>in</strong> a uniform gravitational field Assume<br />

that the rocket ejects mass at a constant rate of ˙ = − where is a positive constant. Solve the equation of<br />

motion to derive the dependence of velocity on time.<br />

c) The first couple of m<strong>in</strong>utes of the launch of the Space Shuttle can be described roughly by; <strong>in</strong>itial mass<br />

= 2 × 10 6 kg, mass after 2 m<strong>in</strong>utes = 1 × 10 6 kg, exhaust speed = 3000 and <strong>in</strong>itial velocity is zero.<br />

Estimate the velocity of the Space Shuttle after two m<strong>in</strong>utes of flight.<br />

d) Describe what would happen to a rocket where ˙ <br />

12. A time <strong>in</strong>dependent field is conservative if ∇ × =0. Use this fact to test if the follow<strong>in</strong>g fields are<br />

conservative, and derive the correspond<strong>in</strong>g potential .<br />

a) = + + = + = + <br />

b) = − − =ln = − +


50 CHAPTER 2. REVIEW OF NEWTONIAN MECHANICS<br />

13. Consider a solid cyl<strong>in</strong>der of mass and radius slid<strong>in</strong>g without roll<strong>in</strong>g down the smooth <strong>in</strong>cl<strong>in</strong>ed face of a<br />

wedge of mass that is free to slide without friction on a horizontal plane floor. Use the coord<strong>in</strong>ates shown<br />

<strong>in</strong> the figure.<br />

a) How far has the wedge moved by the time the cyl<strong>in</strong>der has descended from rest a vertical distance ?<br />

b) Now suppose that the cyl<strong>in</strong>der is free to roll down the wedge without slipp<strong>in</strong>g. How far does the wedge<br />

move <strong>in</strong> this case if the cyl<strong>in</strong>der rolls down a vertical distance ?<br />

c) In which case does the cyl<strong>in</strong>der reach the bottom faster? How does this depend on the radius of the cyl<strong>in</strong>der?<br />

y<br />

.<br />

x<br />

y<br />

x<br />

14. If the gravitational field vector is <strong>in</strong>dependent of the radial distance with<strong>in</strong> a sphere, f<strong>in</strong>d the function describ<strong>in</strong>g<br />

the mass density () of the sphere.


Chapter 3<br />

L<strong>in</strong>ear oscillators<br />

3.1 Introduction<br />

Oscillations are a ubiquitous feature <strong>in</strong> nature. Examples are periodic motion of planets, the rise and fall<br />

of the tides, water waves, pendulum <strong>in</strong> a clock, musical <strong>in</strong>struments, sound waves, electromagnetic waves,<br />

and wave-particle duality <strong>in</strong> quantal physics. Oscillatory systems all have the same basic mathematical form<br />

although the names of the variables and parameters are different. The classical l<strong>in</strong>ear theory of oscillations<br />

will be assumed <strong>in</strong> this chapter s<strong>in</strong>ce: (1) The l<strong>in</strong>ear approximation is well obeyed when the amplitudes of<br />

oscillation are small, that is, the restor<strong>in</strong>g force obeys Hooke’s Law. (2) The Pr<strong>in</strong>ciple of Superposition<br />

applies. (3) The l<strong>in</strong>ear theory allows most problems to be solved explicitly <strong>in</strong> closed form. This is <strong>in</strong> contrast<br />

to non-l<strong>in</strong>ear system where the motion can be complicated and even chaotic as discussed <strong>in</strong> chapter 4.<br />

3.2 L<strong>in</strong>ear restor<strong>in</strong>g forces<br />

An oscillatory system requires that there be a stable equilibrium about<br />

which the oscillations occur. Consider a conservative system with potential<br />

energy for which the force is given by<br />

F = −∇ (3.1)<br />

Figure 31 illustrates a conservative system that has three locations at<br />

which the restor<strong>in</strong>g force is zero, that is, where the gradient of the potential<br />

is zero. Stable oscillations occur only around locations 1 and 3 whereas<br />

the system is unstable at the zero gradient location 2. Po<strong>in</strong>t 2 is called a<br />

separatrix <strong>in</strong> that an <strong>in</strong>f<strong>in</strong>itessimal displacement of the particle from this<br />

separatrix will cause the particle to diverge towards either m<strong>in</strong>imum 1 or<br />

3 depend<strong>in</strong>g on which side of the separatrix the particle is displaced.<br />

The requirements for stable oscillations about any po<strong>in</strong>t 0 are that<br />

the potential energy must have the follow<strong>in</strong>g properties.<br />

Figure 3.1: Stability for a onedimensional<br />

potential U(x).<br />

Stability requirements<br />

1) The potential has a stable position for which the restor<strong>in</strong>g force is zero, i.e. ¡ ¢<br />

<br />

= 0<br />

=0<br />

³<br />

<br />

2) The potential must be positive and an even function of displacement − 0 That is.<br />

<br />

<br />

0<br />

´ 0<br />

where is even.<br />

The requirement for the restor<strong>in</strong>g force to be l<strong>in</strong>ear is that the restor<strong>in</strong>g force for perturbation about a<br />

stable equilibrium at 0 is of the form<br />

F = −(− 0 )=¨ (3.2)<br />

The potential energy function for a l<strong>in</strong>ear oscillator has a pure parabolic shape about the m<strong>in</strong>imum location,<br />

that is,<br />

= 1 2 ( − 0) 2 (3.3)<br />

51


52 CHAPTER 3. LINEAR OSCILLATORS<br />

where 0 is the location of the m<strong>in</strong>imum.<br />

Fortunately, oscillatory systems <strong>in</strong>volve small amplitude oscillations about a stable m<strong>in</strong>imum. For weak<br />

non-l<strong>in</strong>ear systems, where the amplitude of oscillation ∆ about the m<strong>in</strong>imum is small, it is useful to make<br />

a Taylor expansion of the potential energy about the m<strong>in</strong>imum. That is<br />

(∆) =( 0 )+∆ ( 0)<br />

<br />

By def<strong>in</strong>ition, at the m<strong>in</strong>imum (0)<br />

<br />

∆ = (∆) − ( 0 )= ∆2<br />

2!<br />

+ ∆2<br />

2!<br />

2 ( 0 )<br />

2<br />

+ ∆3<br />

3!<br />

3 ( 0 )<br />

3<br />

+ ∆4<br />

4!<br />

=0 and thus equation 33 can be written as<br />

2 ( 0 )<br />

2<br />

+ ∆3<br />

3!<br />

3 ( 0 )<br />

3<br />

+ ∆4<br />

4!<br />

4 ( 0 )<br />

4 + (3.4)<br />

4 ( 0 )<br />

4 + (3.5)<br />

2 ( 0)<br />

2<br />

For small amplitude oscillations, the system is l<strong>in</strong>ear if the second-order ∆2<br />

2!<br />

term <strong>in</strong> equation 32 is<br />

dom<strong>in</strong>ant.<br />

The l<strong>in</strong>earity for small amplitude oscillations greatly simplifies description of the oscillatory motion and<br />

complicated chaotic motion is avoided. Most physical systems are approximately l<strong>in</strong>ear for small amplitude<br />

oscillations, and thus the motion close to equilibrium approximates a l<strong>in</strong>ear harmonic oscillator.<br />

3.3 L<strong>in</strong>earity and superposition<br />

An important aspect of l<strong>in</strong>ear systems is that the solutions obey the Pr<strong>in</strong>ciple of Superposition, thatis,for<br />

the superposition of different oscillatory modes, the amplitudes add l<strong>in</strong>early. The l<strong>in</strong>early-damped l<strong>in</strong>ear<br />

oscillator is an example of a l<strong>in</strong>ear system <strong>in</strong> that it <strong>in</strong>volves only l<strong>in</strong>ear operators, that is, it can be written<br />

<strong>in</strong> the operator form (appendix 2)<br />

µ <br />

2<br />

2 + Γ <br />

+ 2 () = cos (3.6)<br />

The quantity <strong>in</strong> the brackets on the left hand side is a l<strong>in</strong>ear operator that can be designated by L where<br />

L() = () (3.7)<br />

An important feature of l<strong>in</strong>ear operators is that they obey the pr<strong>in</strong>ciple of superposition. This property<br />

results from the fact that l<strong>in</strong>ear operators are distributive, that is<br />

L( 1 + 2 )=L ( 1 )+L ( 2 ) (3.8)<br />

Therefore if there are two solutions 1 () and 2 () for two different forc<strong>in</strong>g functions 1 () and 2 ()<br />

L 1 () = 1 () (3.9)<br />

L 2 () = 2 ()<br />

then the addition of these two solutions, with arbitrary constants, also is a solution for l<strong>in</strong>ear operators.<br />

In general then<br />

L( 1 1 + 2 2 )= 1 1 ()+ 2 2 () (3.10)<br />

à <br />

! Ã<br />

X<br />

<br />

!<br />

X<br />

L () = ()<br />

=1<br />

=1<br />

(3.11)<br />

The left hand bracket can be identified as the l<strong>in</strong>ear comb<strong>in</strong>ation of solutions<br />

X<br />

() = () (3.12)<br />

=1<br />

while the driv<strong>in</strong>g force is a l<strong>in</strong>ear superposition of harmonic forces<br />

() =<br />

X<br />

() (3.13)<br />

=1


3.4. GEOMETRICAL REPRESENTATIONS OF DYNAMICAL MOTION 53<br />

Thus these l<strong>in</strong>ear comb<strong>in</strong>ations also satisfy the general l<strong>in</strong>ear equation<br />

L() = () (3.14)<br />

Applicability of the Pr<strong>in</strong>ciple of Superposition to a system provides a tremendous advantage for handl<strong>in</strong>g<br />

and solv<strong>in</strong>g the equations of motion of oscillatory systems.<br />

3.4 Geometrical representations of dynamical motion<br />

The powerful pattern-recognition capabilities of the human bra<strong>in</strong>, coupled with geometrical representations<br />

of the motion of dynamical systems, provide a sensitive probe of periodic motion. The geometry of the<br />

motion often can provide more <strong>in</strong>sight <strong>in</strong>to the dynamics than <strong>in</strong>spection of mathematical functions. A<br />

system with degrees of freedom is characterized by locations , velocities ˙ and momenta <strong>in</strong> addition<br />

to the time and <strong>in</strong>stantaneous energy (). Geometrical representations of the dynamical correlations are<br />

illustrated by the configuration space and phase space representations of these 2 +2variables.<br />

3.4.1 Configuration space ( )<br />

Aconfiguration space plot shows the correlated motion of two spatial coord<strong>in</strong>ates and averaged over<br />

time. An example is the two-dimensional l<strong>in</strong>ear oscillator with two equations of motion and solutions<br />

q<br />

<br />

¨ + =0 ¨ + =0 (3.15)<br />

() = cos ( ) () = cos ( − ) (3.16)<br />

where =<br />

. For unequal restor<strong>in</strong>g force constants, 6= the trajectory executes complicated<br />

Lissajous figures that depend on the angular frequencies and the phase factor . When the ratio of<br />

the angular frequencies along the two axes is rational, that is <br />

<br />

is a rational fraction, then the curve will<br />

repeat at regular <strong>in</strong>tervals as shown <strong>in</strong> figure 32 and this shape depends on the phase difference. Otherwise<br />

the trajectory uniformly traverses the whole rectangle.<br />

Figure 3.2: Configuration plots of ( ) where =cos(4) and =cos(5 − ) at four different phase values<br />

. The curves are called Lissajous figures


54 CHAPTER 3. LINEAR OSCILLATORS<br />

3.4.2 State space, ( ˙ )<br />

Visualization of a trajectory is enhanced by correlation of configuration and it’s correspond<strong>in</strong>g velocity<br />

˙ which specifies the direction of the motion. The state space representation 1 is especially valuable when<br />

discuss<strong>in</strong>g Lagrangian mechanics which is based on the Lagrangian (q ˙q).<br />

The free undamped harmonic oscillator provides a simple illustration of state space. Consider a mass <br />

attached to a spr<strong>in</strong>g with l<strong>in</strong>ear spr<strong>in</strong>g constant for which the equation of motion is<br />

− = ¨ = ˙ ˙<br />

(3.17)<br />

<br />

By <strong>in</strong>tegration this gives<br />

1<br />

2 ˙2 + 1 2 2 = (3.18)<br />

The first term <strong>in</strong> equation 318 is the k<strong>in</strong>etic energy, the second term is the potential energy, and is the<br />

total energy which is conserved for this system. This equation can be expressed <strong>in</strong> terms of the state space<br />

coord<strong>in</strong>ates as<br />

˙ 2<br />

¡ 2<br />

<br />

¢ + ¡ 2<br />

2<br />

¢ =1 (3.19)<br />

<br />

This corresponds to the equation of an ellipse for a state-space plot of ˙ versus as shown <strong>in</strong> figure 33.<br />

The elliptical paths shown correspond to contours of constant total energy which is partitioned between<br />

k<strong>in</strong>etic and potential energy. For the coord<strong>in</strong>ate axis shown, the motion of a representative po<strong>in</strong>t will be <strong>in</strong><br />

a clockwise direction as the total oscillator energy is redistributed between potential to k<strong>in</strong>etic energy. The<br />

area of the ellipse is proportional to the total energy .<br />

3.4.3 Phase space, ( )<br />

Phase space, which was <strong>in</strong>troduced by J.W. Gibbs for the field of statistical<br />

mechanics, provides a fundamental graphical representation <strong>in</strong><br />

classical mechanics. The phase space coord<strong>in</strong>ates are the conjugate<br />

coord<strong>in</strong>ates (q p) and are fundamental to Hamiltonian mechanics<br />

which is based on the Hamiltonian (q p). For a conservative system,<br />

only one phase-space curve passes through any po<strong>in</strong>t <strong>in</strong> phase space<br />

like the flow of an <strong>in</strong>compressible fluid. This makes phase space more<br />

useful than state space where many curves pass through any location.<br />

Lanczos [La49] def<strong>in</strong>ed an extended phase space us<strong>in</strong>g four-dimensional<br />

relativistic space-time as discussed <strong>in</strong> chapter 17.<br />

S<strong>in</strong>ce = ˙ for the non-relativistic, one-dimensional, l<strong>in</strong>ear oscillator,<br />

then equation 319 can be rewritten <strong>in</strong> the form<br />

2 <br />

2 + ¡ 2<br />

2<br />

¢ =1 (3.20)<br />

<br />

This is the equation of an ellipse <strong>in</strong> the phase space diagram shown <strong>in</strong><br />

Fig.33- which looks identical to Fig 33- where the ord<strong>in</strong>ate<br />

variable = ˙. That is, the only difference is the phase-space coord<strong>in</strong>ates<br />

( ) replace the state-space coord<strong>in</strong>ates ( ˙). State space<br />

plots are used extensively <strong>in</strong> this chapter to describe oscillatory motion.<br />

Although phase space is more fundamental, both state space and<br />

phase space plots provide useful representations for characteriz<strong>in</strong>g and<br />

elucidat<strong>in</strong>g a wide variety of motion <strong>in</strong> classical mechanics. The follow<strong>in</strong>g<br />

discussion of the undamped simple pendulum illustrates the general<br />

features of state space.<br />

Figure 3.3: State space (upper),<br />

and phase space (lower) diagrams,<br />

for the l<strong>in</strong>ear harmonic oscillator.<br />

1 A universal name for the (q ˙q) representation has not been adopted <strong>in</strong> the literature. Therefore this book has adopted<br />

the name "state space" <strong>in</strong> common with reference [Ta05]. Lanczos [La49] uses the term "state space" to refer to the extended<br />

phase space (q p) discussed <strong>in</strong> chapter 17


3.4. GEOMETRICAL REPRESENTATIONS OF DYNAMICAL MOTION 55<br />

3.4.4 Plane pendulum<br />

Consider a simple plane pendulum of mass attachedtoastr<strong>in</strong>goflength <strong>in</strong> a uniform gravitational field<br />

. There is only one generalized coord<strong>in</strong>ate, S<strong>in</strong>ce the moment of <strong>in</strong>ertia of the simple plane-pendulum is<br />

= 2 then the k<strong>in</strong>etic energy is<br />

= 1 2 2 ˙2<br />

(3.21)<br />

and the potential energy relative to the bottom dead center is<br />

= (1 − cos ) (3.22)<br />

Thus the total energy equals<br />

= 1 2 2 ˙2 + (1 − cos ) = + (1 − cos ) (3.23)<br />

22 where is a constant of motion. Note that the angular momentum is not a constant of motion s<strong>in</strong>ce the<br />

angular acceleration ˙ explicitly depends on .<br />

³<br />

It is <strong>in</strong>terest<strong>in</strong>g to look at the solutions for the equation of motion for a plane pendulum on a ˙<br />

´<br />

state space diagram shown <strong>in</strong> figure 34. The curves shown are equally-spaced contours of constant total<br />

energy. Note that the trajectories are ellipses only at very small angles where 1−cos ≈ 2 , the contours are<br />

non-elliptical for higher amplitude oscillations. When the energy is <strong>in</strong> the range 0 2 the motion<br />

corresponds to oscillations of the pendulum about =0. The center of the ellipse is at (0 0) which is a<br />

stable equilibrium po<strong>in</strong>t for the oscillation. However, when || 2 there is a phase change to rotational<br />

motion about the horizontal axis, that is, the pendulum sw<strong>in</strong>gs around and over top dead center, i.e. it<br />

rotates cont<strong>in</strong>uously <strong>in</strong> one direction about the horizontal axis. The phase change occurs at =2 and<br />

is designated by the separatrix trajectory.<br />

Figure 34 shows two cycles for to better illustrate<br />

the cyclic nature of the phase diagram. The closed loops,<br />

shown as f<strong>in</strong>e solid l<strong>in</strong>es, correspond to pendulum oscillations<br />

about =0or 2 for 2. The dashed<br />

l<strong>in</strong>es show roll<strong>in</strong>g motion for cases where the total energy<br />

2. The broad solid l<strong>in</strong>e is the separatrix<br />

that separates the roll<strong>in</strong>g and oscillatory motion. Note<br />

that at the separatrix, the k<strong>in</strong>etic energy and ˙ are zero<br />

when the pendulum is at top dead center which occurs<br />

when = ±The po<strong>in</strong>t ( 0) is an unstable equilibrium<br />

characterized by phase l<strong>in</strong>es that are hyperbolic<br />

to this unstable equilibrium po<strong>in</strong>t. Note that =+<br />

and − correspond to the same physical po<strong>in</strong>t, that is,<br />

the phase diagram is better presented on a cyl<strong>in</strong>drical<br />

phase space representation s<strong>in</strong>ce is a cyclic variable<br />

that cycles around the cyl<strong>in</strong>der whereas ˙ oscillates<br />

equally about zero hav<strong>in</strong>g both positive and negative values.<br />

The state-space diagram can be wrapped around a lum. The axis is <strong>in</strong> units of radians. Note that<br />

Figure 3.4: State space diagram for a plane pendu-<br />

cyl<strong>in</strong>der, then the unstable and stable equilibrium po<strong>in</strong>ts =+ and − correspond to the same physical<br />

will be at diametrically opposite locations on the surface po<strong>in</strong>t, that is the phase diagram should be rolled<br />

of the cyl<strong>in</strong>der at ˙ =0. For small oscillations about <strong>in</strong>to a cyl<strong>in</strong>der connected at = ±.<br />

equilibrium, also called librations, the correlation between<br />

˙ and is given by the clockwise closed loops wrapped on the cyl<strong>in</strong>drical surface, whereas for energies<br />

|| 2 the positive ˙ corresponds to counterclockwise rotations while the negative ˙ corresponds to<br />

clockwise rotations.<br />

State-space diagrams will be used for describ<strong>in</strong>g oscillatory motion <strong>in</strong> chapters 3 and 4 Phase space is<br />

used <strong>in</strong> statistical mechanics <strong>in</strong> order to handle the equations of motion for ensembles of ∼ 10 23 <strong>in</strong>dependent<br />

particles s<strong>in</strong>ce momentum is more fundamental than velocity. Rather than try to account separately for<br />

the motion of each particle for an ensemble, it is best to specify the region of phase space conta<strong>in</strong><strong>in</strong>g the<br />

ensemble. If the number of particles is conserved, then every po<strong>in</strong>t <strong>in</strong> the <strong>in</strong>itial phase space must transform<br />

to correspond<strong>in</strong>g po<strong>in</strong>ts <strong>in</strong> the f<strong>in</strong>al phase space. This will be discussed <strong>in</strong> chapters 83 and 1527.<br />

2


56 CHAPTER 3. LINEAR OSCILLATORS<br />

3.5 L<strong>in</strong>early-damped free l<strong>in</strong>ear oscillator<br />

3.5.1 General solution<br />

All simple harmonic oscillations are damped to some degree due to energy dissipation via friction, viscous<br />

forces, or electrical resistance etc. The motion of damped systems is not conservative <strong>in</strong> that energy is<br />

dissipated as heat. As was discussed <strong>in</strong> chapter 2 the damp<strong>in</strong>g force can be expressed as<br />

F () =−()bv (3.24)<br />

where the velocity dependent function () can be complicated. Fortunately there is a very large class of<br />

problems <strong>in</strong> electricity and magnetism, classical mechanics, molecular, atomic, and nuclear physics, where<br />

the damp<strong>in</strong>g force depends l<strong>in</strong>early on velocity which greatly simplifies solution of the equations of motion.<br />

This chapter discusses the special case of l<strong>in</strong>ear damp<strong>in</strong>g.<br />

Consider the free simple harmonic oscillator, that is, assum<strong>in</strong>g no oscillatory forc<strong>in</strong>g function, with a<br />

l<strong>in</strong>ear damp<strong>in</strong>g term F () =−v where the parameter is the damp<strong>in</strong>g factor. Then the equation of<br />

motion is<br />

− − ˙ = ¨ (3.25)<br />

This can be rewritten as<br />

¨ + Γ ˙ + 2 0 =0 (3.26)<br />

where the damp<strong>in</strong>g parameter<br />

Γ = <br />

(3.27)<br />

and the characteristic angular frequency<br />

0 =<br />

r<br />

<br />

<br />

(3.28)<br />

The general solution to the l<strong>in</strong>early-damped free oscillator is obta<strong>in</strong>ed by <strong>in</strong>sert<strong>in</strong>g the complex trial<br />

solution = 0 Then<br />

() 2 0 + Γ 0 + 2 0 0 =0 (3.29)<br />

This implies that<br />

2 − Γ − 2 0 =0 (3.30)<br />

The solution is<br />

± = Γ 2 ± s<br />

2 0 − µ Γ<br />

2<br />

2<br />

(3.31)<br />

The two solutions ± are complex conjugates and thus the solutions of the damped free oscillator are<br />

= 1 <br />

Γ 2 + <br />

2 0 −( Γ 2 ) 2 <br />

<br />

+ 2 <br />

Γ 2 − <br />

2 0 −( Γ 2 ) 2 <br />

(3.32)<br />

This can be written as<br />

where<br />

= −( Γ 2 ) £ 1 1 + 2 −1¤ (3.33)<br />

s<br />

µ 2 Γ<br />

1 ≡ 2 −<br />

(3.34)<br />

2


3.5. LINEARLY-DAMPED FREE LINEAR OSCILLATOR 57<br />

Underdamped motion 2 1 ≡ 2 − ¡ Γ<br />

2<br />

¢ 2<br />

0<br />

When 2 1 0 then the square root is real so the solution can be written tak<strong>in</strong>g the real part of which<br />

gives that equation 333 equals<br />

() = −( Γ 2 ) cos ( 1 − ) (3.35)<br />

Where and are adjustable constants fit to the <strong>in</strong>itial conditions. Therefore the velocity is given by<br />

˙() =− − Γ 2<br />

∙ 1 s<strong>in</strong> ( 1 − )+ Γ ¸<br />

2 cos ( 1 − )<br />

(3.36)<br />

This is the damped s<strong>in</strong>usoidal oscillation illustrated <strong>in</strong> figure 35.<br />

characteristics:<br />

a) The oscillation amplitude decreases exponentially with a time constant = 2 Γ <br />

The solution has the follow<strong>in</strong>g<br />

q<br />

b) There is a small reduction <strong>in</strong> the frequency of the oscillation due to the damp<strong>in</strong>g lead<strong>in</strong>g to 1 =<br />

2 − ¡ ¢<br />

Γ 2<br />

2<br />

Figure 3.5: The amplitude-time dependence and state-space diagrams for the free l<strong>in</strong>early-damped harmonic<br />

oscillator. The upper row shows the underdamped system for the case with damp<strong>in</strong>g Γ = 0<br />

5<br />

. The lower<br />

row shows the overdamped ( Γ 2 0) [solid l<strong>in</strong>e] and critically damped ( Γ 2 = 0) [dashed l<strong>in</strong>e] <strong>in</strong> both cases<br />

assum<strong>in</strong>g that <strong>in</strong>itially the system is at rest.


58 CHAPTER 3. LINEAR OSCILLATORS<br />

Figure 3.6: Real and imag<strong>in</strong>ary solutions ± of the damped harmonic oscillator. A phase transition occurs<br />

at Γ =2 0 For Γ 2 0 (dashed) the two solutions are complex conjugates and imag<strong>in</strong>ary. For Γ 2 0 ,<br />

(solid), there are two real solutions + and − with widely different decay constants where + dom<strong>in</strong>ates<br />

the decay at long times.<br />

Overdamped case 2 1 ≡ 2 − ¡ Γ<br />

2<br />

¢ 2<br />

0<br />

¡<br />

In this case the square root of 2 1 is imag<strong>in</strong>ary and can be expressed as 0 1 = q<br />

Γ<br />

¢ 2<br />

2 − <br />

2 Therefore the<br />

solution is obta<strong>in</strong>ed more naturally by us<strong>in</strong>g a real trial solution = 0 <strong>in</strong> equation 333 which leads to<br />

two roots<br />

⎡ s ⎤<br />

µΓ<br />

± = − ⎣− Γ 2<br />

2 ± − <br />

2<br />

2 ⎦<br />

<br />

Thus the exponentially damped decay has two time constants + and − <br />

() = £ 1 − + + 2 − − ¤ (3.37)<br />

1<br />

The time constant<br />

−<br />

1<br />

+<br />

thus the first term 1 − + <strong>in</strong> the bracket decays <strong>in</strong> a shorter time than the<br />

second term 2 −− As illustrated <strong>in</strong> figure 36 the decay rate, which is imag<strong>in</strong>ary when underdamped, i.e.<br />

Γ<br />

2 bifurcates <strong>in</strong>to two real values ± for overdamped, i.e. Γ 2 . At large times the dom<strong>in</strong>ant term<br />

when overdamped is for + which has the smallest decay rate, that is, the longest decay constant + = 1<br />

+<br />

.<br />

There is no oscillatory motion for the overdamped case, it slowly moves monotonically to zero as shown <strong>in</strong><br />

fig 35. The amplitude decays away with a time constant that is longer than 2 Γ <br />

Critically damped 2 1 ≡ 2 − ¡ ¢<br />

Γ 2<br />

2 =0<br />

This is the limit<strong>in</strong>g case where Γ 2 = For this case the solution is of the form<br />

() =( + ) −( Γ 2 )<br />

(3.38)<br />

This motion also is non-s<strong>in</strong>usoidal and evolves monotonically to zero. As shown <strong>in</strong> figure 35 the criticallydamped<br />

solution goes to zero with the shortest time constant, that is, largest . Thus analog electric meters<br />

are built almost critically damped so the needle moves to the new equilibrium value <strong>in</strong> the shortest time<br />

without oscillation.<br />

It is useful to graphically represent the motion of the damped l<strong>in</strong>ear oscillator on either a state space<br />

( ˙ ) diagram or phase space ( ) diagram as discussed <strong>in</strong> chapter 34. The state space plots for the<br />

undamped, overdamped, and critically-damped solutions of the damped harmonic oscillator are shown <strong>in</strong><br />

figure 35 For underdamped motion the state space diagram spirals <strong>in</strong>wards to the orig<strong>in</strong> <strong>in</strong> contrast to<br />

critical or overdamped motion where the state and phase space diagrams move monotonically to zero.


3.5. LINEARLY-DAMPED FREE LINEAR OSCILLATOR 59<br />

3.5.2 Energy dissipation<br />

The <strong>in</strong>stantaneous energy is the sum of the <strong>in</strong>stantaneous k<strong>in</strong>etic and potential energies<br />

where and ˙ are given by the solution of the equation of motion.<br />

Consider the total energy of the underdamped system<br />

= 1 2 ˙2 + 1 2 2 (3.39)<br />

= 1 2 2 + 1 2 2 0 2 (3.40)<br />

where = 2 0 The average total energy is given by substitution for and ˙ and tak<strong>in</strong>g the average over<br />

one cycle. S<strong>in</strong>ce<br />

() = −( Γ 2 ) cos ( 1 − ) (3.41)<br />

Then the velocity is given by<br />

˙() =− − Γ 2<br />

∙ 1 s<strong>in</strong> ( 1 − )+ Γ ¸<br />

2 cos ( 1 − )<br />

(3.42)<br />

Insert<strong>in</strong>g equations 341 and 342 <strong>in</strong>to 340 gives a small amplitude oscillation aboutDan exponential decay for<br />

the energy . Averag<strong>in</strong>g over one cycle and us<strong>in</strong>g the fact that hs<strong>in</strong> cos i =0,and [s<strong>in</strong> ] 2E D<br />

= [cos ] 2E =<br />

1<br />

2<br />

, gives the time-averaged total energy as<br />

Ã<br />

hi = −Γ 1<br />

4 2 2 1 + 1 µ 2 Γ<br />

4 2 + 1 2 4 2 0!<br />

2 (3.43)<br />

which can be written as<br />

hi = 0 −Γ (3.44)<br />

Note that the energy of the l<strong>in</strong>early damped free oscillator decays away with a time constant = 1 Γ That<br />

is, the <strong>in</strong>tensity has a time constant that is half the time constant for the decay of the amplitude of the<br />

transient response. Note that the average k<strong>in</strong>etic and potential energies are identical, as implied by the<br />

Virial theorem, and both decay away with the same time constant. This relation between the mean life <br />

for decay of the damped harmonic oscillator and the damp<strong>in</strong>g width term Γ occurs frequently <strong>in</strong> physics.<br />

The damp<strong>in</strong>g of an oscillator usually is characterized by a s<strong>in</strong>gle parameter called the Quality Factor<br />

where<br />

Energy stored <strong>in</strong> the oscillator<br />

≡ (3.45)<br />

Energy dissipated per radian<br />

The energy loss per radian is given by<br />

∆ = <br />

<br />

1<br />

1<br />

= Γ<br />

1<br />

=<br />

q<br />

Γ<br />

2 − ¡ ¢<br />

(3.46)<br />

Γ 2<br />

2<br />

q<br />

where the numerator 1 = 2 − ¡ ¢<br />

Γ 2<br />

2 is the frequency of the free damped l<strong>in</strong>ear oscillator.<br />

Thus the Quality factor equals<br />

.<br />

= ∆ = 1<br />

(3.47)<br />

Γ<br />

The larger the factor, the less damped is the system, and the<br />

greater is the number of cycles of the oscillation <strong>in</strong> the damped<br />

wave tra<strong>in</strong>. Chapter 3113 shows that the longer the wave tra<strong>in</strong>,<br />

that is the higher is the factor, the narrower is the frequency<br />

distribution around the central value. The Mössbauer effect <strong>in</strong><br />

nuclear physics provides a remarkably long wave tra<strong>in</strong> that can<br />

be used to make high precision measurements. The high- precision<br />

of the LIGO laser <strong>in</strong>terferometer was used <strong>in</strong> the first successful<br />

observation of gravity waves <strong>in</strong> 2015.<br />

Typical Q factors<br />

Earth, for earthquake wave 250-1400<br />

Piano str<strong>in</strong>g 3000<br />

Crystal <strong>in</strong> digital watch 10 4<br />

Microwave cavity 10 4<br />

Excited atom 10 7<br />

Neutron star 10 12<br />

LIGO laser 10 13<br />

Mössbauer effect <strong>in</strong> nucleus 10 14<br />

Table 3.1: Typical Q factors <strong>in</strong> nature.


60 CHAPTER 3. LINEAR OSCILLATORS<br />

3.6 S<strong>in</strong>usoidally-drive, l<strong>in</strong>early-damped, l<strong>in</strong>ear oscillator<br />

The l<strong>in</strong>early-damped l<strong>in</strong>ear oscillator, driven by a harmonic driv<strong>in</strong>g force, is of considerable importance to<br />

all branches of science and eng<strong>in</strong>eer<strong>in</strong>g. The equation of motion can be written as<br />

¨ + Γ ˙ + 2 0 = ()<br />

<br />

(3.48)<br />

where () is the driv<strong>in</strong>g force. For mathematical simplicity the driv<strong>in</strong>g force is chosen to be a s<strong>in</strong>usoidal<br />

harmonic force. The solution of this second-order differential equation comprises two components, the<br />

complementary solution (transient response), and the particular solution (steady-state response).<br />

3.6.1 Transient response of a driven oscillator<br />

The transient response of a driven oscillator is given by the complementary solution of the above second-order<br />

differential equation<br />

¨ + Γ ˙ + 2 0 =0 (3.49)<br />

which is identical to the solution of the free l<strong>in</strong>early-damped harmonic oscillator. As discussed <strong>in</strong> section 35<br />

the solution of the l<strong>in</strong>early-damped free oscillator is given by the real part of the complex variable where<br />

= − Γ 2 £ 1 1 + 2 −1¤ (3.50)<br />

and<br />

s<br />

µ 2 Γ<br />

1 ≡ 2 −<br />

(3.51)<br />

2<br />

Underdamped motion 2 1 ≡ 2 − Γ 2<br />

2<br />

0: When <br />

2<br />

1 0 then the square root is real so the transient<br />

solution can be written tak<strong>in</strong>g the real part of which gives<br />

() = 0<br />

− Γ 2 cos ( 1 ) (3.52)<br />

The solution has the follow<strong>in</strong>g characteristics:<br />

a) The amplitude of the transient solution decreases exponentially with a time constant = 2 Γ while<br />

the energy decreases with a time constant of 1 Γ <br />

b) There is a small downward frequency shift <strong>in</strong> that 1 =<br />

q<br />

2 − ¡ Γ<br />

2<br />

¢ 2<br />

Overdamped case 2 1 ≡ 2 − ¡ ¢<br />

Γ 2<br />

2 0: Inthiscasethesquarerootisimag<strong>in</strong>ary,whichcanbeexpressed<br />

q ¡<br />

as 0 1 ≡ Γ<br />

¢ 2<br />

2 − <br />

2 which is real and the solution is just an exponentially damped one<br />

() = h 0<br />

− Γ 2 0 1 + −0 i 1<br />

(3.53)<br />

There is no oscillatory motion for the overdamped case, it slowly moves monotonically to zero. The total<br />

energy decays away with two time constants greater than 1 Γ <br />

Critically damped 2 1 ≡ 2 − ¡ Γ<br />

2<br />

¢ 2<br />

=0: For this case, as mentioned for the damped free oscillator, the<br />

solution is of the form<br />

() =( + ) − Γ 2 (3.54)<br />

The critically-damped system has the shortest time constant.


3.6. SINUSOIDALLY-DRIVE, LINEARLY-DAMPED, LINEAR OSCILLATOR 61<br />

3.6.2 Steady state response of a driven oscillator<br />

The particular solution of the differential equation gives the important steady state response, () to the<br />

forc<strong>in</strong>g function. Consider that the forc<strong>in</strong>g term is a s<strong>in</strong>gle frequency s<strong>in</strong>usoidal oscillation.<br />

() = 0 cos() (3.55)<br />

Thus the particular solution is the real part of the complex variable which is a solution of<br />

A trial solution is<br />

¨ + Γ ˙ + 2 0 = 0<br />

(3.56)<br />

= 0 (3.57)<br />

This leads to the relation<br />

− 2 0 + Γ 0 + 2 0 0 = 0<br />

<br />

Multiply<strong>in</strong>g the numerator and denom<strong>in</strong>ator by the factor ¡ 2 0 − 2¢ − Γ gives<br />

0<br />

<br />

(3.58)<br />

0<br />

0 =<br />

( 2 0 − 2 )+Γ = £¡<br />

<br />

2<br />

( 2 0 − 2 ) 2 +(Γ) 2 0 − 2¢ − Γ ¤ (3.59)<br />

The steady state solution () thus is given by the real part of , thatis<br />

() <br />

=<br />

0<br />

<br />

( 2 0 − 2 ) 2 +(Γ) 2 £¡<br />

<br />

2<br />

0 − 2¢ cos + Γ s<strong>in</strong> ¤ (3.60)<br />

This can be expressed <strong>in</strong> terms of a phase def<strong>in</strong>ed as<br />

µ Γ<br />

tan ≡<br />

2 0 − 2<br />

(3.61)<br />

As shown <strong>in</strong> figure 37 the hypotenuse of the triangle equals<br />

q<br />

( 2 0 − 2 ) 2 +(Γ) 2 .Thus<br />

cos =<br />

2 0 − 2<br />

q( 2 0 − 2 ) 2 +(Γ) 2 (3.62)<br />

and<br />

s<strong>in</strong> =<br />

Γ<br />

q( 2 0 − 2 ) 2 +(Γ) 2 (3.63)<br />

The phase represents the phase difference between the<br />

driv<strong>in</strong>g force and the resultant motion. For a fixed 0 the<br />

phase = 0 when = 0 and <strong>in</strong>creases to = 2<br />

when<br />

= 0 .For 0 the phase → as →∞.<br />

The steady state solution can be re-expressed <strong>in</strong> terms of<br />

the phase shift as<br />

Figure 3.7: Phase between driv<strong>in</strong>g force and<br />

resultant motion.<br />

() <br />

=<br />

0<br />

<br />

q( 2 0 − 2 ) 2 +(Γ) 2 [cos cos +s<strong>in</strong> s<strong>in</strong> ]<br />

=<br />

0<br />

<br />

q( 2 0 − 2 ) 2 +(Γ) 2 cos ( − ) (3.64)


62 CHAPTER 3. LINEAR OSCILLATORS<br />

Figure 3.8: Amplitude versus time, and state space plots of the transient solution (dashed) and total solution<br />

(solid) for two cases. The upper row shows the case where the driv<strong>in</strong>g frequency = 1<br />

5<br />

while the lower row<br />

shows the same for the case where the driv<strong>in</strong>g frequency =5 1 <br />

3.6.3 Complete solution of the driven oscillator<br />

To summarize, the total solution of the s<strong>in</strong>usoidally forced l<strong>in</strong>early-damped harmonic oscillator is the sum<br />

of the transient and steady-state solutions of the equations of motion.<br />

() = () + () (3.65)<br />

For the underdamped case, the transient solution is the complementary solution<br />

where 1 =<br />

() = 0<br />

− Γ 2 cos ( 1 − ) (3.66)<br />

q<br />

2 − ¡ ¢<br />

Γ 2.<br />

2 The steady-state solution is given by the particular solution<br />

() =<br />

0<br />

<br />

q( 2 0 − 2 ) 2 +(Γ) 2 cos ( − ) (3.67)<br />

Note that the frequency of the transient solution is 1 which <strong>in</strong> general differs from the driv<strong>in</strong>g frequency<br />

. The phase shift − for the transient component is set by the <strong>in</strong>itial conditions. The transient response<br />

leads to a more complicated motion immediately after the driv<strong>in</strong>g function is switched on. Figure 38<br />

illustrates the amplitude time dependence and state space diagram for the transient component, and the<br />

total response, when the driv<strong>in</strong>g frequency is either = 1<br />

5<br />

or =5 1 Note that the modulation of the<br />

steady-state response by the transient response is unimportant once the transient response has damped out<br />

lead<strong>in</strong>g to a constant elliptical state space trajectory. For cases where the <strong>in</strong>itial conditions are = ˙ =0<br />

then the transient solution has a relative phase difference − = radians at =0and relative amplitudes<br />

such that the transient and steady-state solutions cancel at =0<br />

The characteristic sounds of different types of musical <strong>in</strong>struments depend very much on the admixture<br />

of transient solutions plus the number and mixture of oscillatory active modes. Percussive <strong>in</strong>struments, such<br />

as the piano, have a large transient component. The mixture of transient and steady-state solutions for<br />

forced oscillations occurs frequently <strong>in</strong> studies of networks <strong>in</strong> electrical circuit analysis.


3.6. SINUSOIDALLY-DRIVE, LINEARLY-DAMPED, LINEAR OSCILLATOR 63<br />

3.6.4 Resonance<br />

The discussion so far has discussed the role of the transient and steady-state solutions of the driven damped<br />

harmonic oscillator which occurs frequently is science, and eng<strong>in</strong>eer<strong>in</strong>g. Another important aspect is resonance<br />

that occurs when the driv<strong>in</strong>g frequency approaches the natural frequency 1 of the damped system.<br />

Considerthecasewherethetimeissufficient for the transient solution to have decayed to zero.<br />

Figure 39 shows the amplitude and phase for the steadystate<br />

response as goes through a resonance as the driv<strong>in</strong>g<br />

frequency is changed. The steady-states solution of the<br />

driven oscillator follows the driv<strong>in</strong>g force when 0 <strong>in</strong><br />

that the phase difference is zero and the amplitude is just<br />

0<br />

<br />

The response of the system peaks at resonance, while<br />

for 0 the harmonic system is unable to follow the<br />

more rapidly oscillat<strong>in</strong>g driv<strong>in</strong>g force and thus the phase of<br />

the <strong>in</strong>duced oscillation is out of phase with the driv<strong>in</strong>g force<br />

and the amplitude of the oscillation tends to zero.<br />

Note that the resonance frequency for a driven damped<br />

oscillator, differs from that for the undriven damped oscillator,<br />

and differs from that for the undamped oscillator. The<br />

natural frequency for an undamped harmonic oscillator<br />

is given by<br />

2 0 = (3.68)<br />

<br />

The transient solution is the same as damped free oscillations<br />

of a damped oscillator and has a frequency of<br />

the system 1 given by<br />

2 1 = 2 0 − µ Γ<br />

2<br />

2<br />

(3.69)<br />

That is, damp<strong>in</strong>g slightly reduces the frequency.<br />

For the driven oscillator the maximum value of the<br />

steady-state amplitude response is obta<strong>in</strong>ed by tak<strong>in</strong>g the<br />

maximum of the function () , that is when <br />

<br />

=0 This<br />

occurs at the resonance angular frequency where<br />

µ 2 Γ<br />

2 = 2 0 − 2<br />

(3.70)<br />

2<br />

Figure 3.9: Resonance behavior for the<br />

l<strong>in</strong>early-damped, harmonically driven, l<strong>in</strong>ear<br />

oscillator.<br />

No resonance occurs if 2 0−2 ¡ Γ<br />

2<br />

¢ 2<br />

0 s<strong>in</strong>ce then is imag<strong>in</strong>ary and the amplitude decreases monotonically<br />

with <strong>in</strong>creas<strong>in</strong>g Note that the above three frequencies are identical if Γ =0but they differ when Γ 0<br />

and 1 0 <br />

For the driven oscillator it is customary to def<strong>in</strong>e the quality factor as<br />

≡ <br />

(3.71)<br />

Γ<br />

When 1 the system has a narrow high resonance peak. As the damp<strong>in</strong>g <strong>in</strong>creases the quality factor<br />

decreases lead<strong>in</strong>g to a wider and lower peak. The resonance disappears when 1 .<br />

3.6.5 Energy absorption<br />

Discussion of energy stored <strong>in</strong> resonant systems is best described us<strong>in</strong>g the steady state solution which is<br />

dom<strong>in</strong>ant after the transient solution has decayed to zero. Then<br />

() <br />

=<br />

0<br />

<br />

( 2 0 − 2 ) 2 +(Γ) 2 £¡<br />

<br />

2<br />

0 − 2¢ cos + Γ s<strong>in</strong> ¤ (3.72)


64 CHAPTER 3. LINEAR OSCILLATORS<br />

This can be rewritten as<br />

where the elastic amplitude<br />

while the absorptive amplitude<br />

() = cos + s<strong>in</strong> (3.73)<br />

=<br />

0<br />

<br />

( 2 0 − 2 ) 2 +(Γ) 2 ¡<br />

<br />

2<br />

0 − 2¢ (3.74)<br />

=<br />

Figure 310 shows the behavior of the absorptive and<br />

elastic amplitudes as a function of angular frequency .<br />

The absorptive amplitude is significant only near resonance<br />

whereas the elastic amplitude goes to zero at<br />

resonance. Note that the full width at half maximum of<br />

the absorptive amplitude peak equals Γ<br />

The work done by the force 0 cos on the oscillator<br />

is<br />

Z Z<br />

= = ˙ (3.76)<br />

Thus the absorbed power () is given by<br />

0<br />

<br />

( 2 0 − 2 ) 2 Γ (3.75)<br />

2<br />

+(Γ)<br />

() = <br />

= ˙ (3.77)<br />

The steady state response gives a velocity<br />

˙() = − s<strong>in</strong> + cos (3.78)<br />

Thus the steady-state <strong>in</strong>stantaneous power <strong>in</strong>put is<br />

() = 0 cos [− s<strong>in</strong> + cos ] (3.79)<br />

Figure 3.10: Elastic (solid) and absorptive<br />

(dashed) amplitudes of the steady-state solution<br />

for Γ =010 0 <br />

The absorptive term steadily absorbs energy while the elastic term oscillates as energy is alternately absorbed<br />

or emitted. The time average over one cycle is given by<br />

h i = 0<br />

h<br />

− hcos s<strong>in</strong> i + <br />

D(cos ) 2Ei (3.80)<br />

where hcos s<strong>in</strong> i and ­ cos 2® are the time average over one cycle. The time averages over one complete<br />

cycle for the first term <strong>in</strong> the bracket is<br />

− hcos s<strong>in</strong> i =0 (3.81)<br />

while for the second term<br />

­<br />

cos <br />

2 ® = 1 <br />

Z 0+<br />

<br />

cos 2 = 1 2<br />

Thus the time average power <strong>in</strong>put is determ<strong>in</strong>ed by only the absorptive term<br />

(3.82)<br />

h i = 1 2 0 = 0<br />

2 Γ 2<br />

2 ( 2 0 − 2 ) 2 +(Γ) 2 (3.83)<br />

This shape of the power curve is a classic Lorentzian shape. Note that the maximum of the average k<strong>in</strong>etic<br />

energy occurs at = 0 which is different from the peak of the amplitude which occurs at 2 1 = 2 0 −¡ ¢<br />

Γ 2.<br />

2<br />

The potential energy is proportional to the amplitude squared, i.e. 2 which occurs at the same angular<br />

frequency as the amplitude, that is, 2 = 2 = 2 0 − 2 ¡ ¢<br />

Γ 2.<br />

2 The k<strong>in</strong>etic and potential energies resonate<br />

at different angular frequencies as a result of the fact that the driven damped oscillator is not conservative


3.6. SINUSOIDALLY-DRIVE, LINEARLY-DAMPED, LINEAR OSCILLATOR 65<br />

because energy is cont<strong>in</strong>ually exchanged between the oscillator and the driv<strong>in</strong>g force system <strong>in</strong> addition to<br />

the energy dissipation due to the damp<strong>in</strong>g.<br />

When ∼ 0 Γ, then the power equation simplifies s<strong>in</strong>ce<br />

¡<br />

<br />

2<br />

0 − 2¢ =( 0 + )( 0 − ) ≈ 2 0 ( 0 − ) (3.84)<br />

Therefore<br />

h i' 0<br />

2<br />

8<br />

Γ<br />

( 0 − ) 2 + ¡ Γ<br />

2<br />

¢ 2<br />

(3.85)<br />

This is called the Lorentzian or Breit-Wigner shape. The half power po<strong>in</strong>ts are at a frequency difference<br />

from resonance of ±∆ where<br />

∆ = | 0 − | = ± Γ (3.86)<br />

2<br />

Thus the full width at half maximum of the Lorentzian curve equals Γ Note that the Lorentzian has a<br />

narrower peak but much wider tail relative to a Gaussian shape. At the peak of the absorbed power, the<br />

absorptive amplitude can be written as<br />

( = 0 )= 0<br />

<br />

<br />

2 0<br />

(3.87)<br />

That is, the peak amplitude <strong>in</strong>creases with <strong>in</strong>crease <strong>in</strong> . This expla<strong>in</strong>s the classic comedy scene where the<br />

soprano shatters the crystal glass because the highest quality crystal glass has a high which leads to a<br />

large amplitude oscillation when she s<strong>in</strong>gs on resonance.<br />

The mean lifetime of the free l<strong>in</strong>early-damped harmonic oscillator, that is, the time for the energy of<br />

free oscillations to decay to 1 was shown to be related to the damp<strong>in</strong>g coefficient Γ by<br />

= 1 Γ<br />

(3.88)<br />

Therefore we have the classical uncerta<strong>in</strong>ty pr<strong>in</strong>ciple for the l<strong>in</strong>early-damped harmonic oscillator<br />

that the measured full-width at half maximum of the energy resonance curve for forced oscillation and the<br />

mean life for decay of the energy of a free l<strong>in</strong>early-damped oscillator are related by<br />

Γ =1 (3.89)<br />

This relation is correct only for a l<strong>in</strong>early-damped harmonic system. Comparable relations between the<br />

lifetime and damp<strong>in</strong>g width exist for different forms of damp<strong>in</strong>g.<br />

One can demonstrate the above l<strong>in</strong>e width and decay time relationship us<strong>in</strong>g an acoustically driven<br />

electric guitar str<strong>in</strong>g. Similarily, the width of the electromagnetic radiation is related to the lifetime for<br />

decay of atomic or nuclear electromagnetic decay. This classical uncerta<strong>in</strong>ty pr<strong>in</strong>ciple is exactly the same<br />

as the one encountered <strong>in</strong> quantum physics due to wave-particle duality. In nuclear physics it is difficult to<br />

measure the lifetime of states when 10 −13 For shorter lifetimes the value of Γ can be determ<strong>in</strong>ed from<br />

the shape of the resonance curve which can be measured directly when the damp<strong>in</strong>g is large.<br />

3.1 Example: Harmonically-driven series RLC circuit<br />

The harmonically-driven, resonant, series circuit, is encountered frequently<br />

<strong>in</strong> AC circuits. Kirchhoff’s Rules applied to the series circuit<br />

lead to the differential equation<br />

¨ + ˙ + = 0 s<strong>in</strong> <br />

where is charge, L is the <strong>in</strong>ductance, is the capacitance, is the resistance,<br />

and the applied voltage across the circuit is () = 0 s<strong>in</strong> . The l<strong>in</strong>earity of<br />

the network allows use of the phasor approach which assumes that the current<br />

= 0 the voltage = 0 (+) and the impedance is a complex number


66 CHAPTER 3. LINEAR OSCILLATORS<br />

= 0<br />

0<br />

where is the phase difference between the voltage and the current. For this circuit the impedance<br />

is given by<br />

µ<br />

= + − 1 <br />

<br />

Because of the phases <strong>in</strong>volved <strong>in</strong> this circuit, at resonance the maximum voltage across the resistor<br />

occurs at a frequency of = 0 across the capacitor the maximum voltage occurs at a frequency 2 =<br />

2 0 − 2<br />

2<br />

and across the <strong>in</strong>ductor the maximum voltage occurs at a frequency 2 2 = 2 0<br />

where 2 1− 2<br />

0 = 1<br />

<br />

2<br />

is the resonance angular frequency when =0. Thus these resonance frequencies differ when 2 0.<br />

3.7 Wave equation<br />

Wave motion is a ubiquitous feature <strong>in</strong> nature. Mechanical wave motion is manifest by transverse waves<br />

on fluid surfaces, longitud<strong>in</strong>al and transverse seismic waves travell<strong>in</strong>g through the Earth, and vibrations of<br />

mechanical structures such as suspended cables. Acoustical wave motion occurs on the stretched str<strong>in</strong>gs of<br />

the viol<strong>in</strong>, as well as the cavities of w<strong>in</strong>d <strong>in</strong>struments. Wave motion occurs for deformable bodies where<br />

elastic forces act<strong>in</strong>g between the nearest-neighbor atoms of the body exert time-dependent forces on one<br />

another. Electromagnetic wave motion <strong>in</strong>cludes wavelengths rang<strong>in</strong>g from 10 5 radiowaves, to 10 −13 rays.<br />

Matter waves are a prom<strong>in</strong>ent feature of quantum physics. All these manifestations of waves exhibit<br />

the same general features of wave motion. Chapter 14 will <strong>in</strong>troduce the collective modes of motion, called<br />

the normal modes, of coupled, many-body, l<strong>in</strong>ear oscillators which act as <strong>in</strong>dependent modes of motion.<br />

The basic elements of wavemotion are <strong>in</strong>troduced at this juncture because the equations of wave motion are<br />

simple, and wave motion features prom<strong>in</strong>ently <strong>in</strong> several chapters throughout this book.<br />

Consider a travell<strong>in</strong>g wave <strong>in</strong> one dimension for a l<strong>in</strong>ear system. If the wave is mov<strong>in</strong>g, then the wave<br />

function Ψ ( ) describ<strong>in</strong>g the shape of the wave, is a function of both and . The <strong>in</strong>stantaneous amplitude<br />

of the wave Ψ ( ) could correspond to the transverse displacement of a wave on a str<strong>in</strong>g, the longitud<strong>in</strong>al<br />

amplitude of a wave on a spr<strong>in</strong>g, the pressure of a longitud<strong>in</strong>al sound wave, the transverse electric or magnetic<br />

fields <strong>in</strong> an electromagnetic wave, a matter wave, etc. If the wave tra<strong>in</strong> ma<strong>in</strong>ta<strong>in</strong>s its shape as it moves, then<br />

one can describe the wave tra<strong>in</strong> by the function () where the coord<strong>in</strong>ate is measured relative to the<br />

shape of the wave, that is, it could correspond to the phase of a crest of the wave. Consider that ( =0)<br />

corresponds to a constant phase, e.g. the peak of the travell<strong>in</strong>g pulse, then assum<strong>in</strong>g that the wave travels<br />

at a phase velocity <strong>in</strong> the direction and the peak is at =0for =0 then it is at = at time .<br />

That is, a po<strong>in</strong>t with phase fixedwithrespecttothewaveformshapeofthewaveprofile () moves <strong>in</strong><br />

the + direction for = − and <strong>in</strong> − direction for = + .<br />

General wave motion can be described by solutions of a wave equation. The wave equation can be<br />

written <strong>in</strong> terms of the spatial and temporal derivatives of the wave function Ψ() Consider the first partial<br />

derivatives of Ψ() =( ∓ ) =()<br />

and<br />

Ψ<br />

= Ψ <br />

= Ψ<br />

<br />

Ψ<br />

= Ψ <br />

<br />

Factor<strong>in</strong>g out Ψ<br />

for the first derivatives gives Ψ Ψ<br />

= ∓<br />

<br />

= ∓<br />

Ψ<br />

<br />

(3.90)<br />

(3.91)<br />

(3.92)<br />

The sign <strong>in</strong> this equation depends on the sign of the wave velocity mak<strong>in</strong>g it not a generally useful formula.<br />

Consider the second derivatives<br />

and<br />

2 Ψ<br />

2<br />

2 Ψ<br />

2<br />

= 2 Ψ<br />

2 <br />

= 2 Ψ<br />

2 (3.93)<br />

=<br />

2 Ψ<br />

2 <br />

=+2 2 Ψ<br />

2 (3.94)


3.8. TRAVELLING AND STANDING WAVE SOLUTIONS OF THE WAVE EQUATION 67<br />

Factor<strong>in</strong>g out 2 Ψ<br />

gives<br />

2<br />

2 Ψ<br />

2 = 1 2 Ψ<br />

2 2 (3.95)<br />

This wave equation <strong>in</strong> one dimension for a l<strong>in</strong>ear system is <strong>in</strong>dependent of the sign of the velocity. There<br />

are an <strong>in</strong>f<strong>in</strong>ite number of possible shapes of waves both travell<strong>in</strong>g and stand<strong>in</strong>g <strong>in</strong> one dimension, all of these<br />

must satisfy this one-dimensional wave equation. The converse is that any function that satisfies this one<br />

dimensional wave equation must be a wave <strong>in</strong> this one dimension.<br />

The Wave Equation <strong>in</strong> three dimensions is<br />

∇ 2 Ψ ≡ 2 Ψ<br />

2 + 2 Ψ<br />

2 + 2 Ψ<br />

2 = 1 2 Ψ<br />

2 2 (3.96)<br />

There are an unlimited number of possible solutions Ψ to this wave equation, any one of which corresponds<br />

toawavemotionwithvelocity.<br />

The Wave Equation is applicable to all manifestations of wave motion, both transverse and longitud<strong>in</strong>al,<br />

for l<strong>in</strong>ear systems. That is, it applies to waves on a str<strong>in</strong>g, water waves, seismic waves, sound waves,<br />

electromagnetic waves, matter waves, etc. If it can be shown that a wave equation can be derived for any<br />

system, discrete or cont<strong>in</strong>uous, then this is equivalent to prov<strong>in</strong>g the existence of waves of any waveform,<br />

frequency, or wavelength travell<strong>in</strong>g with the phase velocity given by the wave equation.[Cra65]<br />

3.8 Travell<strong>in</strong>g and stand<strong>in</strong>g wave solutions of the wave equation<br />

The wave equation can exhibit both travell<strong>in</strong>g and stand<strong>in</strong>g-wave solutions. Consider a one-dimensional<br />

travell<strong>in</strong>g wave with velocity hav<strong>in</strong>g a specific wavenumber ≡ 2 . Then the travell<strong>in</strong>g wave is best<br />

written <strong>in</strong> terms of the phase of the wave as<br />

Ψ( ) =() 2 (∓) = () (∓) (3.97)<br />

where the wave number ≡ 2 <br />

with be<strong>in</strong>g the wave length, and angular frequency ≡ . Thisparticular<br />

solution satisfies the wave equation and corresponds to a travell<strong>in</strong>g wave with phase velocity = <br />

<br />

<strong>in</strong> the<br />

positive or negative direction depend<strong>in</strong>g on whether the sign is negative or positive. Assum<strong>in</strong>g that the<br />

superposition pr<strong>in</strong>ciple applies, then the superposition of these two particular solutions of the wave equation<br />

can be written as<br />

Ψ( ) =()( (−) + (+) )=() ( − + )=2() cos (3.98)<br />

Thus the superposition of two identical s<strong>in</strong>gle wavelength travell<strong>in</strong>g waves propagat<strong>in</strong>g <strong>in</strong> opposite directions<br />

cancorrespondtoastand<strong>in</strong>g wave solution. Note that a stand<strong>in</strong>g wave is identical to a stationary normal<br />

modeofthesystemdiscussed<strong>in</strong>chapter14. This transformation between stand<strong>in</strong>g and travell<strong>in</strong>g waves can<br />

be reversed, that is, the superposition of two stand<strong>in</strong>g waves, i.e. normal modes, can lead to a travell<strong>in</strong>g<br />

wave solution of the wave equation.<br />

Discussion of waveforms is simplified when us<strong>in</strong>g either of the follow<strong>in</strong>g two limits.<br />

1)Thetimedependenceofthewaveformatagivenlocation = 0 which can be expressed us<strong>in</strong>g a<br />

Fourier decomposition, appendix 2, of the time dependence as a function of angular frequency = 0 .<br />

∞X<br />

∞X<br />

Ψ( 0 )= ( 0 0 − 0 ) = ( 0 ) − 0<br />

(3.99)<br />

=−∞<br />

=−∞<br />

2) The spatial dependence of the waveform at a given <strong>in</strong>stant = 0 which can be expressed us<strong>in</strong>g a<br />

Fourier decomposition of the spatial dependence as a function of wavenumber = 0<br />

∞X<br />

∞X<br />

Ψ( 0 )= (0−10) = ( 0 ) 0 (3.100)<br />

=−∞<br />

=−∞<br />

The above is applicable both to discrete, or cont<strong>in</strong>uous l<strong>in</strong>ear oscillator systems, e.g. waves on a str<strong>in</strong>g.<br />

In summary, stationary normal modes of a system are obta<strong>in</strong>ed by a superposition of travell<strong>in</strong>g waves<br />

travell<strong>in</strong>g <strong>in</strong> opposite directions, or equivalently, travell<strong>in</strong>g waves can result from a superposition of stationary<br />

normal modes.


68 CHAPTER 3. LINEAR OSCILLATORS<br />

3.9 Waveform analysis<br />

3.9.1 Harmonic decomposition<br />

As described <strong>in</strong> appendix , when superposition applies, then a<br />

Fourier series decomposition of the form 3101 canbemadeof<br />

any periodic function where<br />

() =<br />

X<br />

cos( 0 + ) (3.101)<br />

=1<br />

A more general Fourier Transform can be made for an aperiodic<br />

function where<br />

Z<br />

() = ()cos( + ()) (3.102)<br />

Any l<strong>in</strong>ear system that is subject to the forc<strong>in</strong>g function ()<br />

has an output that can be expressed as a l<strong>in</strong>ear superposition<br />

of the solutions of the <strong>in</strong>dividual harmonic components of the<br />

forc<strong>in</strong>g function. Fourier analysis of periodic waveforms <strong>in</strong> terms<br />

of harmonic trigonometric functions plays a key role <strong>in</strong> describ<strong>in</strong>g<br />

oscillatory motion <strong>in</strong> classical mechanics and signal process<strong>in</strong>g<br />

for l<strong>in</strong>ear systems. Fourier’s theorem states that any arbitrary<br />

forc<strong>in</strong>g function () canbedecomposed<strong>in</strong>toasumofharmonic<br />

Figure 3.11: The time and frequency representations<br />

of a system exhibit<strong>in</strong>g beats.<br />

terms. As a consequence two equivalent representations can be used to describe signals and waves; the first<br />

is <strong>in</strong> the time doma<strong>in</strong> which describes the time dependence of the signal. The second is <strong>in</strong> the frequency<br />

doma<strong>in</strong> which describes the frequency decomposition of the signal. Fourier analysis relates these equivalent<br />

representations.<br />

For example, the superposition of two equal <strong>in</strong>tensity harmonic<br />

oscillators <strong>in</strong> the time doma<strong>in</strong> is given by<br />

() = cos ( 1 )+ cos ( 2 )<br />

∙µ ¸ ∙µ ¸<br />

1 + 2 1 − 2<br />

= 2 cos<br />

cos<br />

(3.103)<br />

2<br />

2<br />

which leads to the phenomenon of beats as illustrated for both<br />

the time doma<strong>in</strong> and frequency doma<strong>in</strong> <strong>in</strong> figure 311<br />

3.9.2 The free l<strong>in</strong>early-damped l<strong>in</strong>ear oscillator<br />

The response of the free, l<strong>in</strong>early-damped, l<strong>in</strong>ear oscillator is one<br />

of the most frequently encountered waveforms <strong>in</strong> science and thus<br />

it is useful to <strong>in</strong>vestigate the Fourier transform of this waveform.<br />

The waveform amplitude for the underdamped case, shown <strong>in</strong><br />

figure 35 is given by equation (335), thatis<br />

() = − Γ 2 cos ( 1 − ) ≥ 0 (3.104)<br />

() = 0 0 (3.105)<br />

where 2 1 = 2 0 − ¡ Γ<br />

2<br />

¢ 2<br />

and where 0 is the angular frequency of<br />

the undamped system. The Fourier transform is given by<br />

0<br />

£¡<br />

() =<br />

2<br />

( 2 − 2 1 )2 +(Γ) 2 − 2 ¢ ¤<br />

1 − Γ<br />

which is complex and has the famous Lorentz form.<br />

(3.106)<br />

Figure 3.12: The <strong>in</strong>tensity () 2 and<br />

Fourier transform |()| 2 of the free<br />

l<strong>in</strong>early-underdamped harmonic oscillator<br />

with 0 =10and damp<strong>in</strong>g Γ =1.


3.9. WAVEFORM ANALYSIS 69<br />

The<strong>in</strong>tensityofthewavegives<br />

| ()| 2 = 2 −Γ cos 2 ( 1 − ) (3.107)<br />

| ()| 2 =<br />

2 0<br />

( 2 − 2 1 )2 +(Γ) 2 (3.108)<br />

Note that s<strong>in</strong>ce the average over 2 of cos 2 = 1 2 then the average over the cos2 ( 1 − ) term gives the<br />

<strong>in</strong>tensity () = 2<br />

2 −Γ which has a mean lifetime for the decay of = 1 Γ The | ()|2 distribution has the<br />

classic Lorentzian shape, shown <strong>in</strong> figure 312, which has a full width at half-maximum, FWHM, equal to Γ.<br />

Note that () is complex and thus one also can determ<strong>in</strong>e the phase shift which is given by the ratio of<br />

the imag<strong>in</strong>ary to real parts of equation 3105 i.e. tan =<br />

Γ<br />

( 2 − 2 1) .<br />

The mean lifetime of the exponential decay of the <strong>in</strong>tensity can be determ<strong>in</strong>ed either by measur<strong>in</strong>g <br />

from the time dependence, or measur<strong>in</strong>g the FWHM Γ = 1 <br />

of the Fourier transform | ()|2 . In nuclear<br />

and atomic physics excited levels decay by photon emission with the wave form of the free l<strong>in</strong>early-damped,<br />

l<strong>in</strong>ear oscillator. Typically the mean lifetime usually can be measured when & 10 −12 whereas for<br />

shorter lifetimes the radiation width Γ becomes sufficiently large to be measured. Thus the two experimental<br />

approaches are complementary.<br />

3.9.3 Damped l<strong>in</strong>ear oscillator subject to an arbitrary periodic force<br />

Fourier’s theorem states that any arbitrary forc<strong>in</strong>g function () can be decomposed <strong>in</strong>to a sum of harmonic<br />

terms. Consider the response of a damped l<strong>in</strong>ear oscillator to an arbitrary periodic force.<br />

() =<br />

X<br />

0 ( )cos( + ) (3.109)<br />

=0<br />

For each harmonic term the response of a l<strong>in</strong>early-damped l<strong>in</strong>ear oscillator to the forc<strong>in</strong>g function<br />

() = 0 ()cos( ) is given by equation (365 − 67) to be<br />

() = () + () <br />

⎡<br />

⎤<br />

= 0 ( )<br />

⎣ − Γ 2 1<br />

cos ( 1 − )+<br />

cos ( − ) ⎦ (3.110)<br />

<br />

q( 2 0 − 2 ) 2 +(Γ ) 2<br />

The amplitude is obta<strong>in</strong>ed by substitut<strong>in</strong>g <strong>in</strong>to (3110) the derived values 0()<br />

<br />

3.2 Example: Vibration isolation<br />

from the Fourier analysis.<br />

Frequently it is desired to isolate <strong>in</strong>strumentation from the<br />

<strong>in</strong>fluence of horizontal and vertical external vibrations that exist<br />

<strong>in</strong> the environment. One arrangement to achieve this isolation<br />

is to mount a heavy base of mass on weak spr<strong>in</strong>gs of spr<strong>in</strong>g<br />

constant plus weak damp<strong>in</strong>g. The response of this system is<br />

given by equation 3109 which exhibits a resonance at the angular<br />

frequency 2 = 2 0 − 2 ¡ ¢<br />

Γ 2<br />

2 associated with each resonant<br />

frequency 0 of the system. For each resonant frequency the system<br />

amplifies the vibrational amplitude for angular frequencies<br />

close to resonance that is, below √ 2 0 while it attenuates the Seismic isolation of an optical bench.<br />

vibration roughly by a factor of ¡ 0<br />

¢ 2<br />

<br />

at higher frequencies. To<br />

avoid the amplification near the resonance it is necessary to make 0 very much smaller than the frequency<br />

range of the vibrational spectrum and have a moderately high value. This is achieved by use a very heavy<br />

base and weak spr<strong>in</strong>g constant so that 0 is very small. A typical table may have the resonance frequency<br />

at 05 which is well below typical perturb<strong>in</strong>g vibrational frequencies, and thus the table attenuates the<br />

vibration by 99% at 5 and even more attenuation for higher frequency perturbations. This pr<strong>in</strong>ciple is<br />

used extensively <strong>in</strong> design of vibration-isolation tables for optics or microbalance equipment.


70 CHAPTER 3. LINEAR OSCILLATORS<br />

3.10 Signal process<strong>in</strong>g<br />

It has been shown that the response of the l<strong>in</strong>early-damped l<strong>in</strong>ear oscillator, subject to any arbitrary periodic<br />

force, can be calculated us<strong>in</strong>g a frequency decomposition, (Fourier analysis), of the force, appendix . The<br />

response also can be calculated us<strong>in</strong>g a time-ordered discrete-time sampl<strong>in</strong>g of the pulse shape; that is, the<br />

Green’s function approach, appendix . The l<strong>in</strong>early-damped, l<strong>in</strong>ear oscillator is the simplest example of<br />

a l<strong>in</strong>ear system that exhibits both resonance and frequency-dependent response. Typically physical l<strong>in</strong>ear<br />

systems exhibit far more complicated response functions hav<strong>in</strong>g multiple resonances. For example, an automobile<br />

suspension system <strong>in</strong>volves four wheels and associated spr<strong>in</strong>gs plus dampers allow<strong>in</strong>g the car to<br />

rock sideways, or forward and backward, <strong>in</strong> addition to the up-down motion, when subject to the forces<br />

produced by a rough road. Similarly a suspension bridge or aircraft w<strong>in</strong>g can twist as well as bend due to<br />

air turbulence, or a build<strong>in</strong>g can undergo complicated oscillations due to seismic waves. An acoustic system<br />

exhibits similar complexity. Signal analysis and signal process<strong>in</strong>g is of pivotal importance to elucidat<strong>in</strong>g the<br />

response of complicated l<strong>in</strong>ear systems to complicated periodic forc<strong>in</strong>g functions. Signal process<strong>in</strong>g is used<br />

extensively <strong>in</strong> eng<strong>in</strong>eer<strong>in</strong>g, acoustics, and science.<br />

Theresponseofalow-passfilter, such as an R-C circuit or a coaxial cable, to a <strong>in</strong>put square wave,<br />

shown <strong>in</strong> figure 313, provides a simple example of the relative advantages of us<strong>in</strong>g the complementary<br />

Fourier analysis <strong>in</strong> the frequency doma<strong>in</strong>, or the Green’s discrete-function analysis <strong>in</strong> the time doma<strong>in</strong>. The<br />

response of a repetitive square-wave <strong>in</strong>put signal is shown <strong>in</strong> the time doma<strong>in</strong> plus the Fourier transform to<br />

the frequency doma<strong>in</strong>. The middle curves show the time dependence for the response of the low-pass filter<br />

to an impulse () and the correspond<strong>in</strong>g Fourier transform (). The output of the low-pass filter can<br />

be calculated by fold<strong>in</strong>g the <strong>in</strong>put square wave and impulse time dependence <strong>in</strong> the time doma<strong>in</strong> as shown<br />

on the left or by fold<strong>in</strong>g of their Fourier transforms shown on the right. Work<strong>in</strong>g <strong>in</strong> the frequency doma<strong>in</strong><br />

the response of l<strong>in</strong>ear mechanical systems, such as an automobile suspension or a musical <strong>in</strong>strument, as<br />

well as l<strong>in</strong>ear electronic signal process<strong>in</strong>g systems such as amplifiers, loudspeakers and microphones, can<br />

be treated as black boxes hav<strong>in</strong>g a certa<strong>in</strong> transfer function () describ<strong>in</strong>g the ga<strong>in</strong> and phase shift<br />

versus frequency. That is, the output wave frequency decomposition is<br />

() = ( ) · () (3.111)<br />

Work<strong>in</strong>g <strong>in</strong> the time doma<strong>in</strong>, the the low-pass system has an impulse response () = − ,whichisthe<br />

Fourier transform of the transfer function ( ). In the time doma<strong>in</strong><br />

() =<br />

Z ∞<br />

−∞<br />

() · ( − ) (3.112)<br />

This is shown schematically <strong>in</strong> figure 313. The Fourier transformation connects the three quantities <strong>in</strong> the<br />

time doma<strong>in</strong> with the correspond<strong>in</strong>g three <strong>in</strong> the frequency doma<strong>in</strong>. For example, the impulse response of<br />

the low-pass filter has a fall time of which is related by a Fourier transform to the width of the transfer<br />

function. Thus the time and frequency doma<strong>in</strong> approaches are closely related and give the same result for<br />

the output signal for the low-pass filter to the applied square-wave <strong>in</strong>put signal. The result is that the<br />

higher-frequency components are attenuated lead<strong>in</strong>g to slow rise and fall times <strong>in</strong> the time doma<strong>in</strong>.<br />

Analog signal process<strong>in</strong>g and Fourier analysis were the primary tools to analyze and process all forms of<br />

periodic motion dur<strong>in</strong>g the 20 century. For example, musical <strong>in</strong>struments, mechanical systems, electronic<br />

circuits, all employed resonant systems to enhance the desired frequencies and suppress the undesirable<br />

frequencies and the signals could be observed us<strong>in</strong>g analog oscilloscopes. The remarkable development of<br />

comput<strong>in</strong>g has enabled use of digital signal process<strong>in</strong>g lead<strong>in</strong>g to a revolution <strong>in</strong> signal process<strong>in</strong>g that has<br />

had a profound impact on both science and eng<strong>in</strong>eer<strong>in</strong>g. The digital oscilloscope, which can sample at frequencies<br />

above 10 9 has replaced the analog oscilloscope because it allows sophisticated analysis of each<br />

<strong>in</strong>dividual signal that was not possible us<strong>in</strong>g analog signal process<strong>in</strong>g. For example, the analog approach <strong>in</strong><br />

nuclear physics used t<strong>in</strong>y analog electric signals, produced by many <strong>in</strong>dividual radiation detectors, that were<br />

transmitted hundreds of meters via carefully shielded and expensive coaxial cables to the data room where<br />

the signals were amplified and signal processed us<strong>in</strong>g analog filters to maximize the signal to noise <strong>in</strong> order to<br />

separate the signal from the background noise. Stray electromagnetic radiation picked up via the cables significantly<br />

degraded the signals. The performance and limitations of the analog electronics severely restricted<br />

the pulse process<strong>in</strong>g capabilities. Digital signal process<strong>in</strong>g has rapidly replaced analog signal process<strong>in</strong>g.


3.11. WAVE PROPAGATION 71<br />

Figure 3.13: Response of an electrical circuit to an <strong>in</strong>put square wave. The upper row shows the time<br />

and the exponential-form frequency representations of the square-wave <strong>in</strong>put signal. The middle row gives<br />

the impulse response, and correspond<strong>in</strong>g transfer function for the circuit. The bottom row shows the<br />

correspond<strong>in</strong>g output properties <strong>in</strong> both the time and frequency doma<strong>in</strong>s<br />

Analog to digital detector circuits are built directly <strong>in</strong>to the electronics for each <strong>in</strong>dividual detector so that<br />

only digital <strong>in</strong>formation needs to be transmitted from each detector to the analysis computers. Computer<br />

process<strong>in</strong>g provides unlimited and flexible process<strong>in</strong>g capabilities for the digital signals greatly enhanc<strong>in</strong>g<br />

the response and sensitivity of our detector systems. Digital CD and DVD disks are common application of<br />

digital signal process<strong>in</strong>g.<br />

3.11 Wave propagation<br />

Wave motion typically <strong>in</strong>volves a packet of waves encompass<strong>in</strong>g a f<strong>in</strong>ite number of wave cycles. Information<br />

<strong>in</strong> a wave only can be transmitted by start<strong>in</strong>g, stopp<strong>in</strong>g, or modulat<strong>in</strong>g the amplitude of a wave tra<strong>in</strong>, which<br />

is equivalent to form<strong>in</strong>g a wave packet. For example, a musician will play a note for a f<strong>in</strong>ite time, and this<br />

wave tra<strong>in</strong> propagates out as a wave packet of f<strong>in</strong>ite length. You have no <strong>in</strong>formation as to the frequency<br />

and amplitude of the sound prior to the wave packet reach<strong>in</strong>g you, or after the wave packet has passed you.<br />

The velocity of the wavelets conta<strong>in</strong>ed with<strong>in</strong> the wave packet is called the phase velocity. Foradispersive<br />

system the phase velocity of the wavelets conta<strong>in</strong>ed with<strong>in</strong> the wave packet is frequency dependent and the<br />

shape of the wave packet travels at the group velocity which usually differs from the phase velocity. If<br />

the shape of the wave packet is time dependent, then neither the phase velocity, which is the velocity of the<br />

wavelets, nor the group velocity, which is the velocity of an <strong>in</strong>stantaneous po<strong>in</strong>t fixed to the shape of the<br />

wave packet envelope, represent the actual velocity of the overall wavepacket.<br />

A third wavepacket velocity, the signal velocity, isdef<strong>in</strong>ed to be the velocity of the lead<strong>in</strong>g edge of the<br />

energy distribution, and correspond<strong>in</strong>g <strong>in</strong>formation content, of the wave packet. For most l<strong>in</strong>ear systems<br />

the shape of the wave packet is not time dependent and then the group and signal velocities are identical.<br />

However, the group and signal velocities can be very different for non-l<strong>in</strong>ear systems as discussed <strong>in</strong> chapter<br />

47. Note that even when the phase velocity of the waves with<strong>in</strong> the wave packet travels faster than the group<br />

velocity of the shape, or the signal velocity of the energy content of the envelope of the wave packet, the<br />

<strong>in</strong>formation conta<strong>in</strong>ed <strong>in</strong> a wave packet is only manifest when the wave packet envelope reaches the detector<br />

and this energy and <strong>in</strong>formation travel at the signal velocity. The modern ideas of wave propagation,<br />

<strong>in</strong>clud<strong>in</strong>g Hamilton’s concept of group velocity, were developed by Lord Rayleigh when applied to the theory<br />

of sound[Ray1887]. The concept of phase, group, and signal velocities played a major role <strong>in</strong> discussion of<br />

electromagnetic waves as well as de Broglie’s development of wave-particle duality <strong>in</strong> quantum mechanics.


72 CHAPTER 3. LINEAR OSCILLATORS<br />

3.11.1 Phase, group, and signal velocities of wave packets<br />

The concepts of wave packets, as well as their phase, group, and signal velocities, are of considerable importance<br />

for propagation of <strong>in</strong>formation and other manifestations of wave motion <strong>in</strong> science and eng<strong>in</strong>eer<strong>in</strong>g.<br />

This importance warrants further discussion at this juncture.<br />

Consider a particular component of a one-dimensional wave,<br />

The argument of the exponential is called the phase of the wave where<br />

( ) = (±) (3.113)<br />

≡ − (3.114)<br />

If we move along the axis at a velocity such that the phase is constant then we perceive a stationary<br />

pattern <strong>in</strong> this mov<strong>in</strong>g frame. The velocity of this wave is called the phase velocity. To ensure constant<br />

phase requires that is constant, or assum<strong>in</strong>g real and <br />

Therefore the phase velocity is def<strong>in</strong>ed to be<br />

= (3.115)<br />

= <br />

(3.116)<br />

The velocity discussed so far is just the phase velocity of the <strong>in</strong>dividual wavelets at the carrier frequency. If<br />

or are complex then one must take the real parts to ensure that the velocity is real.<br />

If the phase velocity of a wave is dependent on the wavelength, that is, () then the system is<br />

said to be dispersive <strong>in</strong> that the wave is dispersed accord<strong>in</strong>g the wavelength. The simplest illustration of<br />

dispersion is the refraction of light <strong>in</strong> glass prism which leads to dispersion of the light <strong>in</strong>to the spectrum of<br />

wavelengths. Dispersion leads to development of wave packets that travel at group and signal velocities that<br />

usually differ from the phase velocity. To illustrate this behavior, consider two equal amplitude travell<strong>in</strong>g<br />

waves hav<strong>in</strong>g slightly different wave number and angular frequency . Superposition of these waves gives<br />

( ) = ( [−] + [(+∆)−(+∆)] ) (3.117)<br />

=<br />

∆<br />

∆<br />

[(+<br />

2 )−(+ 2 )] ∆ ∆<br />

−[ ·{ 2 − 2 ] ∆ ∆<br />

[<br />

+ 2 − 2 ] }<br />

=<br />

∆<br />

∆<br />

[(+<br />

2 2 )−(+ 2 )] cos[ ∆<br />

2 − ∆<br />

2 ]<br />

This corresponds to a wave with the average carrier frequency modulated by the cos<strong>in</strong>e term which has a<br />

wavenumber of ∆<br />

2<br />

and angular frequency ∆<br />

2<br />

, that is, this is the usual example of beats The cos<strong>in</strong>e term<br />

modulates the average wave produc<strong>in</strong>g wave packets as shown <strong>in</strong> figure 311. The velocity of these wave<br />

packetsiscalledthegroup velocity given by requir<strong>in</strong>g that the phase of the modulat<strong>in</strong>g term is constant,<br />

that is<br />

∆<br />

2<br />

∆<br />

= (3.118)<br />

2<br />

Thus the group velocity is given by<br />

= <br />

= ∆<br />

(3.119)<br />

∆<br />

If dispersion is present then the group velocity = ∆<br />

∆ does not equal the phase velocity = <br />

Expand<strong>in</strong>g the above example to superposition of waves gives<br />

( ) =<br />

X<br />

(±) (3.120)<br />

=1<br />

In the event that →∞and the frequencies are cont<strong>in</strong>uously distributed, then the summation is replaced<br />

by an <strong>in</strong>tegral<br />

( ) =<br />

Z ∞<br />

−∞<br />

() (±) (3.121)


3.11. WAVE PROPAGATION 73<br />

where the factor () represents the distribution amplitudes of the component waves, that is the spectral<br />

decomposition of the wave. This is the usual Fourier decomposition of the spatial distribution of the wave.<br />

Consider an extension of the l<strong>in</strong>ear superposition of two waves to a well def<strong>in</strong>ed wave packet where the<br />

amplitude is nonzero only for a small range of wavenumbers 0 ± ∆<br />

( ) =<br />

Z 0 +∆<br />

0−∆<br />

() (−) (3.122)<br />

This functional shape is called a wave packet which only has mean<strong>in</strong>g if ∆ 0 . The angular frequency<br />

can be expressed by mak<strong>in</strong>g a Taylor expansion around 0<br />

µ <br />

() =( 0 )+ ( − 0 )+ (3.123)<br />

<br />

0<br />

For a l<strong>in</strong>ear system the phase then reduces to<br />

µ <br />

− =( 0 − 0 )+( − 0 ) − ( − 0 ) (3.124)<br />

<br />

0<br />

The summation of terms <strong>in</strong> the exponent given by 3124 leads to the amplitude 3122 hav<strong>in</strong>gtheformofa<br />

product where the <strong>in</strong>tegral becomes<br />

( ) = (0−0) Z 0 +∆<br />

0−∆<br />

() (− 0)[−( <br />

) 0<br />

]<br />

(3.125)<br />

The <strong>in</strong>tegral term modulates the (0−0) first term.<br />

The group velocity is def<strong>in</strong>ed to be that for which the phase of the exponential term <strong>in</strong> the <strong>in</strong>tegral is<br />

constant. Thus<br />

µ <br />

=<br />

(3.126)<br />

<br />

0<br />

S<strong>in</strong>ce = then<br />

= + <br />

(3.127)<br />

<br />

For non-dispersive systems the phase velocity is <strong>in</strong>dependent of the wave number or angular frequency <br />

and thus = The case discussed earlier, equation (3103) for beat<strong>in</strong>g of two waves gives the<br />

same relation <strong>in</strong> the limit that ∆ and ∆ are <strong>in</strong>f<strong>in</strong>itessimal.<br />

¡ The group velocity of a wave packet is of physical significance for dispersive media where =<br />

<br />

¢<br />

0<br />

6= = . Every wave tra<strong>in</strong> has a f<strong>in</strong>ite extent and thus we usually observe the motion of a<br />

group of waves rather than the wavelets mov<strong>in</strong>g with<strong>in</strong> the wave packet. In general, for non-l<strong>in</strong>ear dispersive<br />

systems the derivative <br />

<br />

can be either positive or negative and thus <strong>in</strong> pr<strong>in</strong>ciple the group velocity<br />

can either be greater than, or less than, the phase velocity. Moreover, if the group velocity is frequency<br />

dependent, that is, when group velocity dispersion occurs, then the overall shape of the wave packet is time<br />

dependent and thus the speed of a specific relativelocationdef<strong>in</strong>ed by the shape of the envelope of the wave<br />

packet does not represent the signal velocity of the wave packet. Brillou<strong>in</strong> showed that the distribution<br />

of the energy, and correspond<strong>in</strong>g <strong>in</strong>formation content, for any wave packet, travels at the signal velocity<br />

which can be different from the group velocity if the shape of the envelope of the wave packet is time<br />

dependent. For electromagnetic waves one has the possibility that the group velocity = In<br />

1914 Brillou<strong>in</strong>[Bri14][Bri60] showed that the signal velocity of electromagnetic waves, def<strong>in</strong>ed by the lead<strong>in</strong>g<br />

edge of the time-dependent envelope of the wave packet, never exceeds even though the group velocity<br />

correspond<strong>in</strong>g to the velocity of the <strong>in</strong>stantaneous shape of the wave packet may exceed . Thus, there is<br />

no violation of E<strong>in</strong>ste<strong>in</strong>’s fundamental pr<strong>in</strong>ciple of relativity that the velocity of an electromagnetic wave<br />

cannot exceed .


74 CHAPTER 3. LINEAR OSCILLATORS<br />

3.3 Example: Water waves break<strong>in</strong>g on a beach<br />

The concepts of phase and group velocity are illustrated by the example of water waves mov<strong>in</strong>g at velocity<br />

<strong>in</strong>cident upon a straight beach at an angle to the shorel<strong>in</strong>e. Consider that the wavepacket comprises<br />

many wavelengths of wavelength . Dur<strong>in</strong>g the time it takes the wave to travel a distance the po<strong>in</strong>t where<br />

the crest of one wave breaks on the beach travels a distance<br />

crest of the one wavelet <strong>in</strong> the wave packet is<br />

=<br />

The velocity of the wave packet along the beach equals<br />

<br />

cos <br />

= cos <br />

<br />

cos <br />

along beach. Thus the phase velocity of the<br />

Note that for the wave mov<strong>in</strong>g parallel to the beach =0and = = . However, for = 2<br />

→∞and → 0. In general for waves break<strong>in</strong>g on the beach<br />

= 2<br />

The same behavior is exhibited by surface waves bounc<strong>in</strong>g off the sides of the Erie canal, sound waves <strong>in</strong><br />

a trombone, and electromagnetic waves transmitted down a rectangular wave guide. In the latter case the<br />

phase velocity exceeds the velocity of light <strong>in</strong> apparent violation of E<strong>in</strong>ste<strong>in</strong>’s theory of relativity. However,<br />

the <strong>in</strong>formation travels at the signal velocity which is less than .<br />

3.4 Example: Surface waves for deep water<br />

In the “Theory of Sound”[Ray1887] Rayleigh discusses the example of surface waves for water. He derives<br />

a dispersion relation for the phase velocity and wavenumber whicharerelatedtothedensity, depth<br />

, gravity , and surface tension , by<br />

2 = + 3<br />

<br />

tanh()<br />

For deep water where the wavelength is short compared with the depth, that is kl 1 then tanh() → 1<br />

and the dispersion relation is given approximately by<br />

2 = + 3<br />

<br />

For long surface waves for deep water, that is, small , then the gravitational first term <strong>in</strong> the dispersion<br />

relation dom<strong>in</strong>ates and the group velocity is given by<br />

µ <br />

= = 1 2<br />

r <br />

= 1 <br />

2 = <br />

2<br />

That is, the group velocity is half of the phase velocity. Here the wavelets are build<strong>in</strong>g at the back of the wave<br />

packet, progress through the wave packet and dissipate at the front. This can be demonstrated by dropp<strong>in</strong>g a<br />

pebble <strong>in</strong>to a calm lake. It will be seen that the surface disturbance comprises a wave packet mov<strong>in</strong>g outwards<br />

at the group velocity with the <strong>in</strong>dividual waves with<strong>in</strong> the wave packet expand<strong>in</strong>g at twice the group velocity<br />

of the wavepacket, that is, they are created at the <strong>in</strong>ner radius of the wave packet and disappear at the outer<br />

radius of the wave packet.<br />

For small wavelength ripples, where is large, then the surface tension term dom<strong>in</strong>ates and the dispersion<br />

relation is approximately given by<br />

2 ' 3<br />

<br />

lead<strong>in</strong>g to a group velocity of<br />

µ <br />

= = 3 2 <br />

Here the group velocity exceeds the phase velocity and wavelets are build<strong>in</strong>g at the front of the wave packet and<br />

dissipate at the back. Note that for this l<strong>in</strong>ear system, the Brillion signal velocity equals the group velocity<br />

for both gravity and surface tension waves for deep water.


3.11. WAVE PROPAGATION 75<br />

3.5 Example: Electromagnetic waves <strong>in</strong> ionosphere<br />

The response to radio waves, <strong>in</strong>cident upon a free electron plasma <strong>in</strong> the ionosphere, provides an excellent<br />

example that <strong>in</strong>volves cut-off frequency, complex wavenumber as well as the phase, group, and signal<br />

velocities. Maxwell’s equations give the most general wave equation for electromagnetic waves to be<br />

∇ 2 E − 2 E<br />

2 = j ³<br />

<br />

´<br />

+ ∇·<br />

<br />

<br />

∇ 2 H − 2 H<br />

2 = −∇ × j <br />

where and j are the unbound charge and current densities. The effect of the bound charges and<br />

currents are absorbed <strong>in</strong>to and . Ohm’s Law can be written <strong>in</strong> terms of the electrical conductivity which<br />

is a constant<br />

j =E<br />

Assum<strong>in</strong>g Ohm’s Law plus assum<strong>in</strong>g =0, <strong>in</strong> the plasma gives the relations<br />

∇ 2 E − 2 E<br />

2<br />

∇ 2 H − 2 H<br />

2<br />

− E <br />

− H <br />

The third term <strong>in</strong> both of these wave equations is a damp<strong>in</strong>g term that leads to a damped solution of an<br />

electromagnetic wave <strong>in</strong> a good conductor.<br />

The solution of these damped wave equations can be solved by consider<strong>in</strong>g an <strong>in</strong>cident wave<br />

E = ˆx (−)<br />

Substitut<strong>in</strong>g for E <strong>in</strong> the first damped wave equation gives<br />

That is<br />

− 2 + 2 − =0<br />

∙<br />

2 = 2 1 − ¸<br />

<br />

In general is complex, that is, it has real and imag<strong>in</strong>ary parts that lead to a solution of the form<br />

E = − (−)<br />

The first exponential term is an exponential damp<strong>in</strong>g term while the second exponential term is the oscillat<strong>in</strong>g<br />

term.<br />

Consider that the plasma <strong>in</strong>volves the motion of a bound damped electron, of charge of mass bound<br />

<strong>in</strong> a one dimensional atom or lattice subject to an oscillatory electric field of frequency . Assume that the<br />

electromagnetic wave is travell<strong>in</strong>g <strong>in</strong> the ˆ direction with the transverse electric field <strong>in</strong> the ˆ direction. The<br />

equation of motion of an electron can be written as<br />

= 0<br />

= 0<br />

ẍ + Γẋ + 2 0 = ˆx 0 (−)<br />

where Γ is the damp<strong>in</strong>g factor. The <strong>in</strong>stantaneous displacement of the oscillat<strong>in</strong>g charge equals<br />

x = 1<br />

( 2 0 − 2 )+Γ ˆx 0 (−)<br />

and the velocity is<br />

Thus the <strong>in</strong>stantaneous current density is given by<br />

ẋ = <br />

( 2 0 − 2 )+Γ ˆx 0 (−)<br />

j = ẋ = 2<br />

<br />

<br />

( 2 0 − 2 )+Γ ˆx 0 (−)


76 CHAPTER 3. LINEAR OSCILLATORS<br />

Therefore the electrical conductivity is given by<br />

= 2<br />

<br />

<br />

( 2 0 − 2 )+Γ<br />

Let us consider only unbound charges <strong>in</strong> the plasma, that is let 0 =0. Then the conductivity is given by<br />

= 2<br />

<br />

<br />

Γ − 2<br />

For a low density ionized plasma Γ thus the conductivity is given approximately by<br />

≈− 2<br />

<br />

<br />

S<strong>in</strong>ce is pure imag<strong>in</strong>ary, then j and E have a phase difference of<br />

2<br />

which implies that the average of<br />

the Joule heat<strong>in</strong>g over a complete period is hj · Ei =0 Thus there is no energy loss due to Joule heat<strong>in</strong>g<br />

imply<strong>in</strong>g that the electromagnetic energy is conserved.<br />

Substitution of <strong>in</strong>to the relation for 2<br />

∙<br />

2 = 2 1 − ¸<br />

¸<br />

= 2 <br />

∙1 − 2<br />

<br />

2<br />

Def<strong>in</strong>e the Plasma oscillation frequency to be<br />

r<br />

<br />

2<br />

≡<br />

<br />

then 2 can be written as<br />

∙ ³<br />

2 = 2 <br />

´2¸<br />

1 −<br />

()<br />

<br />

For a low density plasma the dielectric constant ' 1 and the relative permeability ' 1 and thus<br />

= 0 ' 0 and = 0 ' 0 . The velocity of light <strong>in</strong> vacuum = √ 1<br />

0 <br />

. Thus for low density<br />

0<br />

equation can be written as<br />

2 = 2 + 2 2<br />

()<br />

Differentiation of equation with respect to gives 2 <br />

=22 That is, = 2 and the phase<br />

velocity is<br />

r<br />

= 2 + 2 <br />

2<br />

Therearethreecasestoconsider.<br />

1) : For this case<br />

h1 − ¡ <br />

¢ i 2<br />

1 and thus is a pure real number. Therefore the electromagnetic<br />

wave is transmitted with a phase velocity that exceeds while the group velocity is less than<br />

.<br />

2) : For this case<br />

h1 − ¡ <br />

¢ i 2<br />

1 and thus is a pure imag<strong>in</strong>ary number. Therefore the<br />

electromagnetic wave is not transmitted <strong>in</strong> the ionosphere and is attenuated rapidly as −( <br />

) . However,<br />

s<strong>in</strong>ce there are no Joule heat<strong>in</strong>g losses, then the electromagnetic wave must be complete reflected. Thus the<br />

Plasma oscillation frequency serves as a cut-off frequency. For this example the signal and group velocities<br />

are identical.<br />

For the ionosphere = 10 −11 electrons/m 3 ,whichcorrespondstoaPlasmaoscillationfrequencyof<br />

= 2 =3. Thus electromagnetic waves <strong>in</strong> the AM waveband ( 16) are totally reflected by<br />

the ionosphere and bounce repeatedly around the Earth, whereas for VHF frequencies above 3,thewaves<br />

are transmitted and refracted pass<strong>in</strong>g through the atmosphere. Thus light is transmitted by the ionosphere.<br />

By contrast, for a good conductor like silver, the Plasma oscillation frequency is around 10 16 which is<br />

<strong>in</strong> the far ultraviolet part of the spectrum. Thus, all lower frequencies, such as light, are totally reflected<br />

by such a good conductor, whereas X-rays have frequencies above the Plasma oscillation frequency and are<br />

transmitted.


3.11. WAVE PROPAGATION 77<br />

3.11.2 Fourier transform of wave packets<br />

The relation between the time distribution and the correspond<strong>in</strong>g<br />

frequency distribution, or equivalently, the<br />

spatial distribution and the correspond<strong>in</strong>g wave-number<br />

distribution, are of considerable importance <strong>in</strong> discussion<br />

of wave packets and signal process<strong>in</strong>g. It directly<br />

relates to the uncerta<strong>in</strong>ty pr<strong>in</strong>ciple that is a characteristic<br />

of all forms of wave motion. The relation between the<br />

time and correspond<strong>in</strong>g frequency distribution is given<br />

via the Fourier transform discussed <strong>in</strong> appendix . The<br />

follow<strong>in</strong>g are two examples of the Fourier transforms of<br />

typical but rather differentwavepacketshapesthatare<br />

encountered often <strong>in</strong> science and eng<strong>in</strong>eer<strong>in</strong>g.<br />

3.6 Example: Fourier transform of a<br />

Gaussian wave packet:<br />

Assum<strong>in</strong>g that the amplitude of the wave is a<br />

Gaussian wave packet shown <strong>in</strong> the adjacent figure where<br />

() = − (− 0 )2<br />

2 2 <br />

This leads to the Fourier transform<br />

() = √ 2 − 2 2<br />

2 cos ( 0 )<br />

Fourier transform of a Gaussian frequency<br />

distribution.<br />

Note that the wavepacket has a standard deviation for the amplitude of the wavepacket of = 1<br />

<br />

,that<br />

is · =1. The Gaussian wavepacket results <strong>in</strong> the m<strong>in</strong>imum product of the standard deviations of the<br />

frequency and time representations for a wavepacket. This has profound importance for all wave phenomena,<br />

and especially to quantum mechanics. Because matter exhibits wave-like behavior, the above property of wave<br />

packet leads to Heisenberg’s Uncerta<strong>in</strong>ty Pr<strong>in</strong>ciple. For signal process<strong>in</strong>g, it shows that if you truncate a<br />

wavepacket you will broaden the frequency distribution.<br />

3.7 Example: Fourier transform of a rectangular wave packet:<br />

Assume unity amplitude of the frequency distribution between 0 − ∆ ≤ ≤ 0 + ∆ ,thatis,as<strong>in</strong>gle<br />

isolated square pulse of width that is described by the rectangular function Π def<strong>in</strong>ed as<br />

½ 1<br />

Π() =<br />

0<br />

| − 0 | ∆<br />

| − 0 | ∆<br />

Then the Fourier transform us given by<br />

() =<br />

∙ ¸ s<strong>in</strong> ∆<br />

cos 0 <br />

∆<br />

That is, the transform of a rectangular wavepacket gives a cos<strong>in</strong>e wave modulated by an unnormalized<br />

function which is a nice example of a simple wave packet. That is, on the right hand side we have<br />

a wavepacket ∆ = ± 2<br />

∆<br />

wide. Note that the product of the two measures of the widths ∆ · ∆ = ±<br />

Example 2 considers a rectangular<br />

³ ´<br />

pulse of unity amplitude between − 2 ≤ ≤ 2 whichresulted<strong>in</strong>a<br />

s<strong>in</strong> <br />

Fourier transform () = 2<br />

. That is, for a pulse of width ∆ = ± <br />

2<br />

2<br />

the frequency envelope has<br />

the first zero at ∆ = ± <br />

. Note that this is the complementary system to the one considered here which has<br />

∆ · ∆ = ± illustrat<strong>in</strong>g the symmetry of the Fourier transform and its <strong>in</strong>verse.


78 CHAPTER 3. LINEAR OSCILLATORS<br />

3.11.3 Wave-packet Uncerta<strong>in</strong>ty Pr<strong>in</strong>ciple<br />

The Uncerta<strong>in</strong>ty Pr<strong>in</strong>ciple states that wavemotion exhibits a m<strong>in</strong>imum product of the uncerta<strong>in</strong>ty <strong>in</strong> the<br />

simultaneously measured width <strong>in</strong> time of a wave packet, and the distribution width of the frequency decomposition<br />

of this wave packet. This was illustrated by the Fourier transforms of wave packets discussed<br />

above where it was shown the product of the widths is m<strong>in</strong>imized for a Gaussian-shaped wave packet. The<br />

Uncerta<strong>in</strong>ty Pr<strong>in</strong>ciple implies that to make a precise measurement of the frequency of a s<strong>in</strong>usoidal wave<br />

requiresthatthewavepacketbe<strong>in</strong>f<strong>in</strong>itely long. If the duration of the wave packet is reduced then the<br />

frequency distribution broadens. The crucial aspect qneeded for this discussion, is that, for the amplitudes<br />

of any wavepacket, the standard deviations () = h 2 i − hi 2 characteriz<strong>in</strong>g the width of the spectral<br />

distribution <strong>in</strong> the angular frequency doma<strong>in</strong>, (), and the width for the conjugate variable <strong>in</strong> time ()<br />

are related :<br />

() · () > 1<br />

(Relation between amplitude uncerta<strong>in</strong>ties.)<br />

This product of the standard deviations equals unity only for the special case of Gaussian-shaped spectral<br />

distributions, and it is greater than unity for all other shaped spectral distributions.<br />

The <strong>in</strong>tensity of the wave is the square of the amplitude lead<strong>in</strong>g to standard deviation widths for a<br />

Gaussian distribution where () 2 = 1 2 () 2 ,thatis, () = () √<br />

2<br />

. Thus the standard deviations for the<br />

spectral distribution and width of the <strong>in</strong>tensity of the wavepacket are related by:<br />

() · () > 1 2<br />

(Uncerta<strong>in</strong>ty pr<strong>in</strong>ciple for frequency-time <strong>in</strong>tensities)<br />

This states that the uncerta<strong>in</strong>ties with which you can simultaneously measure the time and frequency<br />

for the <strong>in</strong>tensity of a given wavepacket are related. If you try to measure the frequency with<strong>in</strong> a short time<br />

<strong>in</strong>terval () then the uncerta<strong>in</strong>ty <strong>in</strong> the frequency measurement () > 1<br />

2 ()<br />

Accurate measurement<br />

of the frequency requires measurement times that encompass many cycles of oscillation, that is, a long<br />

wavepacket.<br />

Exactly the same relations exist between the spectral distribution as a function of wavenumber and<br />

the correspond<strong>in</strong>g spatial dependence of a wave which are conjugate representations. Thus the spectral<br />

distribution plotted versus is directly related to the amplitude as a function of position ; the spectral<br />

distribution versus is related to the amplitude as a function of ; and the spectral distribution is related<br />

to the spatial dependence on Follow<strong>in</strong>g the same arguments discussed above, the standard deviation,<br />

( ) characteriz<strong>in</strong>g the width of the spectral <strong>in</strong>tensity distribution of , and the standard deviation<br />

() characteriz<strong>in</strong>g the spatial width of the wave packet <strong>in</strong>tensity as a function of are related by the<br />

Uncerta<strong>in</strong>ty Pr<strong>in</strong>ciple for position-wavenumber. Thus <strong>in</strong> summary the temporal and spatial uncerta<strong>in</strong>ty<br />

pr<strong>in</strong>ciples of the <strong>in</strong>tensity of wave motion is,<br />

() · () > 1 2<br />

(3.128)<br />

() · ( ) > 1 2<br />

() · ( ) > 1 2<br />

() · ( ) > 1 2<br />

This applies to all forms of wave motion, be they, sound waves, water waves, electromagnetic waves, or<br />

matter waves.<br />

As discussed <strong>in</strong> chapter 18, the transition to quantum mechanics <strong>in</strong>volves relat<strong>in</strong>g the matter-wave properties<br />

to the energy and momentum of the correspond<strong>in</strong>g particle. That is, <strong>in</strong> the case of matter waves,<br />

multiply<strong>in</strong>g both sides of equation 3129 by ~ and us<strong>in</strong>g the de Broglie relations gives that the particle energy<br />

is related to the angular frequency by = ~ and the particle momentum is related to the wavenumber,<br />

that is −→ p = ~ −→ k . These lead to the Heisenberg Uncerta<strong>in</strong>ty Pr<strong>in</strong>ciple:<br />

() · () > ~ 2<br />

(3.129)<br />

() · ( ) > ~ 2<br />

() · ( ) > ~ 2<br />

() · ( ) > ~ 2<br />

This uncerta<strong>in</strong>ty pr<strong>in</strong>ciple applies equally to the wavefunction of the electron <strong>in</strong> the<br />

hydrogen atom, proton <strong>in</strong> a nucleus, as well as to a wavepacket describ<strong>in</strong>g a particle wave mov<strong>in</strong>g along some


3.11. WAVE PROPAGATION 79<br />

trajectory. This implies that, for a particle of given momentum, the wavefunction is spread out spatially.<br />

Planck’s constant ~ =105410 −34 · =658210 −16 · is extremely small compared with energies and<br />

times encountered <strong>in</strong> normal life, and thus the effects due to the Uncerta<strong>in</strong>ty Pr<strong>in</strong>ciple are not important for<br />

macroscopic dimensions.<br />

Conf<strong>in</strong>ement of a particle, of mass , with<strong>in</strong>±() of a fixed location implies that there is a correspond<strong>in</strong>g<br />

uncerta<strong>in</strong>ty <strong>in</strong> the momentum<br />

( ) ≥ ~<br />

2()<br />

Now the variance <strong>in</strong> momentum p is given by the difference <strong>in</strong> the average of the square<br />

square of the average of hpi 2 .Thatis<br />

(p) 2 =<br />

(3.130)<br />

D<br />

(p · p) 2E ,andthe<br />

D<br />

(p · p) 2E − hpi 2 (3.131)<br />

Assum<strong>in</strong>g a fixed average location implies that hpi =0,then<br />

D<br />

(p · p) 2E µ 2 ~<br />

= () 2 ≥<br />

(3.132)<br />

2()<br />

S<strong>in</strong>ce the k<strong>in</strong>etic energy is given by:<br />

K<strong>in</strong>etic energy = 2<br />

2 ≥ ~ 2<br />

8() 2<br />

(Zero-po<strong>in</strong>t energy)<br />

This zero-po<strong>in</strong>t energy is the m<strong>in</strong>imum k<strong>in</strong>etic energy that a particle of mass can have if conf<strong>in</strong>ed with<strong>in</strong> a<br />

distance ±() This zero-po<strong>in</strong>t energy is a consequence of wave-particle duality and the uncerta<strong>in</strong>ty between<br />

the size and wavenumber for any wave packet. It is a quantal effect <strong>in</strong> that the classical limit has ~ → 0 for<br />

which the zero-po<strong>in</strong>t energy → 0<br />

Insert<strong>in</strong>g numbers for the zero-po<strong>in</strong>t energy gives that an electron conf<strong>in</strong>ed to the radius of the atom,<br />

that is () =10 −10 has a zero-po<strong>in</strong>t k<strong>in</strong>etic energy of ∼ 1 .Conf<strong>in</strong><strong>in</strong>g this electron to 3 × 10 −15 the<br />

size of a nucleus, gives a zero-po<strong>in</strong>t energy of 10 9 (1 ) Conf<strong>in</strong><strong>in</strong>g a proton to the size of the nucleus<br />

gives a zero-po<strong>in</strong>t energy of 05. These values are typical of the level spac<strong>in</strong>g observed <strong>in</strong> atomic and<br />

nuclear physics. If ~ was a large number, then a billiard ball conf<strong>in</strong>ed to a billiard table would be a blur<br />

as it oscillated with the m<strong>in</strong>imum zero-po<strong>in</strong>t k<strong>in</strong>etic energy. The smaller the spatial region that the ball<br />

was conf<strong>in</strong>ed, the larger would be its zero-po<strong>in</strong>t energy and momentum caus<strong>in</strong>g it to rattle back and forth<br />

between the boundaries of the conf<strong>in</strong>ed region. Life would be dramatically different if ~ was a large number.<br />

In summary, Heisenberg’s Uncerta<strong>in</strong>ty Pr<strong>in</strong>ciple is a well-known and crucially important aspect of quantum<br />

physics. What is less well known, is that the Uncerta<strong>in</strong>ty Pr<strong>in</strong>ciple applies for all forms of wave motion,<br />

that is, it is not restricted to matter waves. The follow<strong>in</strong>g three examples illustrate application of the<br />

Uncerta<strong>in</strong>ty Pr<strong>in</strong>ciple to acoustics, the nuclear Mössbauer effect, and quantum mechanics.<br />

3.8 Example: Acoustic wave packet<br />

A viol<strong>in</strong>ist plays the note middle C (261625) with constant <strong>in</strong>tensity for precisely 2 seconds. Us<strong>in</strong>g<br />

the fact that the velocity of sound <strong>in</strong> air is 3432 calculate the follow<strong>in</strong>g:<br />

1) The wavelength of the sound wave <strong>in</strong> air: = 3432261625 = 1312.<br />

2) The length of the wavepacket <strong>in</strong> air: Wavepacket length = 3432 × 2 = 6864<br />

3) The fractional frequency width of the note: S<strong>in</strong>ce the wave packet has a square pulse shape of length<br />

=2, then the Fourier transform is a s<strong>in</strong>c function hav<strong>in</strong>g the first zeros when s<strong>in</strong> <br />

2 =0,thatis,∆ = 1 .<br />

Therefore the fractional width is ∆<br />

= 1<br />

=00019. Note that to achieve a purity of ∆<br />

<br />

=10−6 the viol<strong>in</strong>ist<br />

would have to play the note for 106.<br />

3.9 Example: Gravitational red shift<br />

The Mössbauer effect <strong>in</strong> nuclear physics provides a wave packet that has an exceptionally small fractional<br />

width <strong>in</strong> frequency. For example, the 57 Fe nucleus emits a 144 deexcitation-energy photon which corresponds<br />

to ≈ 2 × 10 25 withadecaytimeof ≈ 10 −7 . Thus the fractional width is ∆<br />

≈ 3 × 10−18 .


80 CHAPTER 3. LINEAR OSCILLATORS<br />

In 1959 Pound and Rebka used this to test E<strong>in</strong>ste<strong>in</strong>’s general theory of relativity by measurement of the<br />

gravitational red shift between the attic and basement of the 225 high physics build<strong>in</strong>g at Harvard. The<br />

magnitude of the predicted relativistic red shift is ∆<br />

<br />

=25 × 15 10− which is what was observed with a<br />

fractional precision of about 1%.<br />

3.10 Example: Quantum baseball<br />

George Gamow, <strong>in</strong> his book ”Mr. Tompk<strong>in</strong>s <strong>in</strong> Wonderland”, describes the strange world that would exist<br />

if ~ was a large number. As an example, consider you play baseball <strong>in</strong> a universe where ~ is a large number.<br />

The pitcher throws a 150 ball 20 to the batter at a speed of 40. For a strike to be thrown, the ball’s<br />

position must be pitched with<strong>in</strong> the 30 radius of the strike zone, that is, it is required that ∆ ≤ 03.<br />

The uncerta<strong>in</strong>ty relation tells us that the transverse velocity of the ball cannot be less than ∆ =<br />

<br />

2∆ The<br />

time of flight of the ball from the mound to batter is =05. Because of the transverse velocity uncerta<strong>in</strong>ty,<br />

∆ theballwilldeviate∆ transversely from the strike zone. This also must not exceed the size of the<br />

strike zone, that is;<br />

∆ = ~<br />

2∆ ≤ 03<br />

(Due to transverse velocity uncerta<strong>in</strong>ty)<br />

Comb<strong>in</strong><strong>in</strong>g both of these requirements gives<br />

~ ≤ 2∆2<br />

<br />

=54 10 −2 · <br />

This is 32 orders of magnitude larger than ~ so quantal effects are negligible. However, if ~ exceeded the<br />

above value, then the pitcher would have difficulty throw<strong>in</strong>g a reliable strike.<br />

3.12 Summary<br />

L<strong>in</strong>ear systems have the feature that the solutions obey the Pr<strong>in</strong>ciple of Superposition, thatis,theamplitudes<br />

add l<strong>in</strong>early for the superposition of different oscillatory modes. Applicability of the Pr<strong>in</strong>ciple of<br />

Superposition to a system provides a tremendous advantage for handl<strong>in</strong>g and solv<strong>in</strong>g the equations of motion<br />

of oscillatory systems.<br />

Geometric representations of the motion of dynamical systems provide sensitive probes of periodic motion.<br />

Configuration space (q q), statespace(q ˙q) and phase space (q p), are powerful geometric<br />

representations that are used extensively for recogniz<strong>in</strong>g periodic motion where q ˙q and p are vectors <strong>in</strong><br />

-dimensional space.<br />

L<strong>in</strong>early-damped free l<strong>in</strong>ear oscillator<br />

the equation<br />

The free l<strong>in</strong>early-damped l<strong>in</strong>ear oscillator is characterized by<br />

¨ + Γ ˙ + 2 0 =0<br />

(326)<br />

The solutions of the l<strong>in</strong>early-damped free l<strong>in</strong>ear oscillator are of the form<br />

s<br />

= −( Γ 2 ) £ 1 1 + 2 − 1 ¤<br />

µ 2 Γ<br />

1 ≡ 2 − (333)<br />

2<br />

The solutions of the l<strong>in</strong>early-damped free l<strong>in</strong>ear oscillator have the follow<strong>in</strong>g characteristic frequencies correspond<strong>in</strong>g<br />

to the three levels of l<strong>in</strong>ear damp<strong>in</strong>g<br />

() = −( q<br />

Γ<br />

2 ) cos ( 1 − ) underdamped 1 = 2 − ¡ ¢<br />

Γ 2<br />

2 0<br />

∙ q ¡<br />

() =[ 1 − + + 2 −− ] overdamped ± = − − Γ 2 ± Γ<br />

¢ 2<br />

2 − <br />

2 ¸<br />

() =( + ) −( q<br />

Γ<br />

2 )<br />

critically damped 1 = 2 − ¡ ¢<br />

Γ 2<br />

2 =0


3.12. SUMMARY 81<br />

The energy dissipation for the l<strong>in</strong>early-damped free l<strong>in</strong>ear oscillator time averaged over one period is<br />

given by<br />

hi = 0 −Γ<br />

(344)<br />

The quality factor characteriz<strong>in</strong>g the damp<strong>in</strong>g of the free oscillator is def<strong>in</strong>ed to be<br />

= ∆ = 1<br />

Γ<br />

where ∆ is the energy dissipated per radian.<br />

(347)<br />

S<strong>in</strong>usoidally-driven, l<strong>in</strong>early-damped, l<strong>in</strong>ear oscillator The l<strong>in</strong>early-damped l<strong>in</strong>ear oscillator, driven<br />

by a harmonic driv<strong>in</strong>g force, is of considerable importance to all branches of physics, and eng<strong>in</strong>eer<strong>in</strong>g. The<br />

equation of motion can be written as<br />

¨ + Γ ˙ + 2 0 = ()<br />

(349)<br />

<br />

where () is the driv<strong>in</strong>g force. The complete solution of this second-order differential equation comprises<br />

two components, the complementary solution (transient response), and the particular solution (steady-state<br />

response). Thatis,<br />

() = () + () <br />

(365)<br />

For the underdamped case, the transient solution is the complementary solution<br />

() = 0<br />

− Γ 2 cos ( 1 − )<br />

and the steady-state solution is given by the particular solution<br />

(366)<br />

() =<br />

0<br />

<br />

q( 2 0 − 2 ) 2 +(Γ) 2 cos ( − ) (367)<br />

Resonance A detailed discussion of resonance and energy absorption for the driven l<strong>in</strong>early-damped l<strong>in</strong>ear<br />

oscillator was given. For resonance of the l<strong>in</strong>early-damped l<strong>in</strong>ear oscillator the maximum amplitudes occur<br />

at the follow<strong>in</strong>g resonant frequencies<br />

Resonant system<br />

Resonant<br />

q<br />

frequency<br />

<br />

undamped free l<strong>in</strong>ear oscillator 0 =<br />

<br />

q<br />

l<strong>in</strong>early-damped free l<strong>in</strong>ear oscillator 1 = 2 0 − ¡ ¢<br />

Γ 2<br />

q 2<br />

driven l<strong>in</strong>early-damped l<strong>in</strong>ear oscillator = 2 0 − 2 ¡ ¢<br />

Γ 2<br />

2<br />

The energy absorption for the steady-state solution for resonance is given by<br />

() = cos + s<strong>in</strong> <br />

(373)<br />

where the elastic amplitude<br />

while the absorptive amplitude<br />

=<br />

0<br />

<br />

( 2 0 − 2 ) 2 +(Γ) 2 ¡<br />

<br />

2<br />

0 − 2¢ (374)<br />

0<br />

<br />

=<br />

( 2 0 − 2 ) 2 Γ (375)<br />

2<br />

+(Γ)<br />

The time average power <strong>in</strong>put is given by only the absorptive term<br />

h i = 1 2 0 = 0<br />

2 Γ 2<br />

2 ( 2 0 − 2 ) 2 +(Γ) 2 (3.133)<br />

This power curve has the classic Lorentzian shape.


82 CHAPTER 3. LINEAR OSCILLATORS<br />

Wave propagation The wave equation was <strong>in</strong>troduced and both travell<strong>in</strong>g and stand<strong>in</strong>g wave solutions<br />

of the wave equation were discussed. Harmonic wave-form analysis, and the complementary time-sampled<br />

wave form analysis techniques, were <strong>in</strong>troduced <strong>in</strong> this chapter and <strong>in</strong> appendix . The relative merits of<br />

Fourier analysis and the digital Green’s function waveform analysis were illustrated for signal process<strong>in</strong>g.<br />

The concepts of phase velocity, group velocity, and signal velocity were <strong>in</strong>troduced. The phase velocity<br />

is given by<br />

= (3117)<br />

<br />

and group velocity<br />

µ <br />

= = + <br />

(3128)<br />

<br />

0<br />

<br />

If the group velocity is frequency dependent then the <strong>in</strong>formation content of a wave packet travels at the<br />

signal velocity which can differ from the group velocity.<br />

The Wave-packet Uncerta<strong>in</strong>ty Pr<strong>in</strong>ciple implies that mak<strong>in</strong>g a precise measurement of the frequency q of a<br />

s<strong>in</strong>usoidal wave requires that the wave packet be <strong>in</strong>f<strong>in</strong>itely long. The standard deviation () = h 2 i − hi 2<br />

characteriz<strong>in</strong>g the width of the amplitude of the wavepacket spectral distribution <strong>in</strong> the angular frequency<br />

doma<strong>in</strong>, (), and the correspond<strong>in</strong>g width <strong>in</strong> time () are related by :<br />

() · () > 1<br />

(Relation between amplitude uncerta<strong>in</strong>ties.)<br />

The standard deviations for the spectral distribution and width of the <strong>in</strong>tensity ofthewavepacketare<br />

related by:<br />

() · () > 1 (3.134)<br />

2<br />

() · ( ) > 1 2<br />

() · ( ) > 1 2<br />

() · ( ) > 1 2<br />

This applies to all forms of wave motion, <strong>in</strong>clud<strong>in</strong>g sound waves, water waves, electromagnetic waves, or<br />

matter waves.


3.12. SUMMARY 83<br />

Problems<br />

1. Consider a simple harmonic oscillator consist<strong>in</strong>g of a mass attached to a spr<strong>in</strong>g of spr<strong>in</strong>g constant . For<br />

this oscillator () = s<strong>in</strong>( 0 − ).<br />

(a) F<strong>in</strong>d an expression for ˙().<br />

(b) Elim<strong>in</strong>ate between () and ˙() to arrive at one equation similar to that for an ellipse.<br />

(c) Rewrite the equation <strong>in</strong> part (b) <strong>in</strong> terms of , ˙, , , and the total energy .<br />

(d) Give a rough sketch of the phase space diagram ( ˙ versus ) for this oscillator. Also, on the same set of<br />

axes, sketch the phase space diagram for a similar oscillator with a total energy that is larger than the<br />

first oscillator.<br />

(e) What direction are the paths that you have sketched? Expla<strong>in</strong> your answer.<br />

(f) Would different trajectories for the same oscillator ever cross paths? Why or why not?<br />

2. Consider a damped, driven oscillator consist<strong>in</strong>g of a mass attachedtoaspr<strong>in</strong>gofspr<strong>in</strong>gconstant.<br />

(a) What is the equation of motion for this system?<br />

(b) Solve the equation <strong>in</strong> part (a). The solution consists of two parts, the complementary solution and the<br />

particular solution. When might it be possible to safely neglect one part of the solution?<br />

(c) What is the difference between amplitude resonance and k<strong>in</strong>etic energy resonance?<br />

(d) How might phase space diagrams look for this type of oscillator? What variables would affect the diagram?<br />

3. A particle of mass is subject to the follow<strong>in</strong>g force<br />

where is a constant.<br />

F = ( 3 − 4 2 +3)ˆx<br />

(a) Determ<strong>in</strong>e the po<strong>in</strong>ts when the particle is <strong>in</strong> equilibrium.<br />

(b) Which of these po<strong>in</strong>ts is stable and which are unstable?<br />

(c) Is the motion bounded or unbounded?<br />

4. A very long cyl<strong>in</strong>drical shell has a mass density that depends upon the radial distance such that () = ,<br />

where is a constant. The <strong>in</strong>ner radius of the shell is and the outer radius is .<br />

(a) Determ<strong>in</strong>e the direction and the magnitude of the gravitational field for all regions of space.<br />

(b) If the gravitational potential is zero at the orig<strong>in</strong>, what is the difference between the gravitational potential<br />

at = and = ?<br />

5. A mass is constra<strong>in</strong>ed to move along one dimension. Two identical spr<strong>in</strong>gs are attached to the mass, one on<br />

each side, and each spr<strong>in</strong>g is <strong>in</strong> turn attached to a wall. Both spr<strong>in</strong>gs have the same spr<strong>in</strong>g constant .<br />

(a) Determ<strong>in</strong>e the frequency of the oscillation, assum<strong>in</strong>g no damp<strong>in</strong>g.<br />

(b) Now consider damp<strong>in</strong>g. It is observed that after oscillations, the amplitude of the oscillation has<br />

dropped to one-half of its <strong>in</strong>itial value. F<strong>in</strong>d an expression for the damp<strong>in</strong>g constant.<br />

(c) How long does it take for the amplitude to decrease to one-quarter of its <strong>in</strong>itial value?<br />

6. Discussthemotionofacont<strong>in</strong>uousstr<strong>in</strong>gwhenpluckedat ½<br />

one third of the length of the str<strong>in</strong>g. That is, the<br />

3<br />

<strong>in</strong>itial condition is ˙( 0) = 0, and ( 0) = 0 ≤ ≤ ¾<br />

<br />

3<br />

3<br />

<br />

2<br />

( − )<br />

3 ≤ ≤ <br />

7. When a particular driv<strong>in</strong>g force is applied to a stretched str<strong>in</strong>g it is observed that the str<strong>in</strong>g vibration <strong>in</strong> purely<br />

of the harmonic. F<strong>in</strong>d the driv<strong>in</strong>g force.


84 CHAPTER 3. LINEAR OSCILLATORS<br />

8. Consider the two-mass system pivoted at its vertex where 6= . It undergoes oscillations of the angle <br />

with respect to the vertical <strong>in</strong> the plane of the triangle.<br />

l<br />

l<br />

M<br />

l<br />

m<br />

(a) Determ<strong>in</strong>e the angular frequency of small oscillations.<br />

(b) Use your result from part (a) to show 2 ≈ <br />

for À .<br />

(c) Show that your result from part (a) agrees with 2 = 00 ( )<br />

<br />

where is the equilibrium angle and is<br />

the moment of <strong>in</strong>ertia.<br />

(d) Assume the system has energy . Setup an <strong>in</strong>tegral that determ<strong>in</strong>es the period of oscillation.<br />

9. An unusual pendulum is made by fix<strong>in</strong>gastr<strong>in</strong>gtoahorizontalcyl<strong>in</strong>derofradius wrapp<strong>in</strong>g the str<strong>in</strong>g<br />

several times around the cyl<strong>in</strong>der, and then ty<strong>in</strong>g a mass to the loose end. In equilibrium the mass hangs a<br />

distance 0 vertically below the edge of the cyl<strong>in</strong>der. F<strong>in</strong>d the potential energy if the pendulum has swung to<br />

an angle from the vertical. Show that for small angles, it can be written <strong>in</strong> the Hooke’s Law form = 1 2 2 .<br />

Comment of the value of <br />

10. Consider the two-dimensional anisotropic oscillator with motion with = and = .<br />

a) Prove that if the ratio of the frequencies is rational (that is, <br />

<br />

= <br />

where and are <strong>in</strong>tegers) then the<br />

motion is periodic. What is the period?<br />

b) Prove that if the same ratio is irrational, the motion never repeats itself.<br />

11. A simple pendulum consists of a mass suspended from a fixed po<strong>in</strong>t by a weight-less, extensionless rod of<br />

length .<br />

a) Obta<strong>in</strong> the equation of motion, and <strong>in</strong> the approximation s<strong>in</strong> ≈ show that the natural frequency is<br />

0 = p <br />

<br />

,where is the gravitational field strength.<br />

b) Discuss the motion <strong>in</strong> the event that the motion takes place <strong>in</strong> a viscous medium with retard<strong>in</strong>g force<br />

2 √ ˙.<br />

12. Derive the expression for the State Space paths of the plane pendulum if the total energy is 2. Note<br />

that this is just the case of a particle mov<strong>in</strong>g <strong>in</strong> a periodic potential () =(1−cos) Sketch the State<br />

Space diagram for both 2 and 2<br />

13. Consider the motion of a driven l<strong>in</strong>early-damped harmonic oscillator after the transient solution has died out,<br />

and suppose that it is be<strong>in</strong>g driven close to resonance, = .<br />

a) Show that the oscillator’s total energy is = 1 2 2 2 .<br />

b) Show that the energy ∆ dissipated dur<strong>in</strong>g one cycle by the damp<strong>in</strong>g force Γ ˙ is Γ 2<br />

14. Two masses m 1 and m 2 slide freely on a horizontal frictionless rail and are connected by a spr<strong>in</strong>g whose force<br />

constant is k. F<strong>in</strong>d the frequency of oscillatory motion for this system.<br />

15. A particle of mass moves under the <strong>in</strong>fluence of a resistive force proportional to velocity and a potential ,<br />

that is .<br />

( ˙) =− ˙ − <br />

<br />

where 0 and () =( 2 − 2 ) 2<br />

a) F<strong>in</strong>d the po<strong>in</strong>ts of stable and unstable equilibrium.<br />

b) F<strong>in</strong>d the solution of the equations of motion for small oscillations around the stable equilibrium po<strong>in</strong>ts<br />

c) Show that as →∞the particle approaches one of the stable equilibrium po<strong>in</strong>ts for most choices of <strong>in</strong>itial<br />

conditions. What are the exceptions? (H<strong>in</strong>t: You can prove this without f<strong>in</strong>d<strong>in</strong>g the solutions explicitly.)


Chapter 4<br />

Nonl<strong>in</strong>ear systems and chaos<br />

4.1 Introduction<br />

In nature only a subset of systems have equations of motion that are l<strong>in</strong>ear. Contrary to the impression<br />

given by the analytic solutions presented <strong>in</strong> undergraduate physics courses, most dynamical systems <strong>in</strong><br />

nature exhibit non-l<strong>in</strong>ear behavior that leads to complicated motion. The solutions of non-l<strong>in</strong>ear equations<br />

usually do not have analytic solutions, superposition does not apply, and they predict phenomena such as<br />

attractors, discont<strong>in</strong>uous period bifurcation, extreme sensitivity to <strong>in</strong>itial conditions, roll<strong>in</strong>g motion, and<br />

chaos. Dur<strong>in</strong>g the past four decades, excit<strong>in</strong>g discoveries have been made <strong>in</strong> classical mechanics that are<br />

associated with the recognition that nonl<strong>in</strong>ear systems can exhibit chaos. Chaotic phenomena have been<br />

observed <strong>in</strong> most fields of science and eng<strong>in</strong>eer<strong>in</strong>g such as, weather patterns, fluid flow, motion of planets <strong>in</strong><br />

the solar system, epidemics, chang<strong>in</strong>g populations of animals, birds and <strong>in</strong>sects, and the motion of electrons<br />

<strong>in</strong> atoms. The complicated dynamical behavior predicted by non-l<strong>in</strong>ear differential equations is not limited<br />

to classical mechanics, rather it is a manifestation of the mathematical properties of the solutions of the<br />

differential equations <strong>in</strong>volved, and thus is generally applicable to solutions of first or second-order nonl<strong>in</strong>ear<br />

differential equations. It is important to understand that the systems discussed <strong>in</strong> this chapter follow<br />

a fully determ<strong>in</strong>istic evolution predicted by the laws of classical mechanics, the evolution for which is based<br />

on the prior history. This behavior is completely different from a random walk where each step is based on a<br />

random process. The complicated motion of determ<strong>in</strong>istic non-l<strong>in</strong>ear systems stems <strong>in</strong> part from sensitivity<br />

to the <strong>in</strong>itial conditions.<br />

The French mathematician Po<strong>in</strong>caré is credited with be<strong>in</strong>g the first to recognize the existence of chaos<br />

dur<strong>in</strong>g his <strong>in</strong>vestigation of the gravitational three-body problem <strong>in</strong> celestial mechanics. At the end of the<br />

n<strong>in</strong>eteenth century Po<strong>in</strong>caré noticed that such systems exhibit high sensitivity to <strong>in</strong>itial conditions characteristic<br />

of chaotic motion, and the existence of nonl<strong>in</strong>earity which is required to produce chaos. Po<strong>in</strong>caré’s work<br />

received little notice, <strong>in</strong> part it was overshadowed by the parallel development of the Theory of Relativity<br />

and quantum mechanics at the start of the 20 century. In addition, solv<strong>in</strong>g nonl<strong>in</strong>ear equations of motion<br />

is difficult, which discouraged work on nonl<strong>in</strong>ear mechanics and chaotic motion. The field blossomed dur<strong>in</strong>g<br />

the 1960 0 when computers became sufficiently powerful to solve the nonl<strong>in</strong>ear equations required to calculate<br />

the long-time histories necessary to document the evolution of chaotic behavior. Laplace, and many other<br />

scientists, believed <strong>in</strong> the determ<strong>in</strong>istic view of nature which assumes that if the position and velocities of<br />

all particles are known, then one can unambiguously predict the future motion us<strong>in</strong>g Newtonian mechanics.<br />

Researchers <strong>in</strong> many fields of science now realize that this “clockwork universe” is <strong>in</strong>valid. That is, know<strong>in</strong>g<br />

thelawsofnaturecanbe<strong>in</strong>sufficient to predict the evolution of nonl<strong>in</strong>ear systems <strong>in</strong> that the time evolution<br />

can be extremely sensitive to the <strong>in</strong>itial conditions even though they follow a completely determ<strong>in</strong>istic<br />

development. There are two major classifications of nonl<strong>in</strong>ear systems that lead to chaos <strong>in</strong> nature. The<br />

first classification encompasses nondissipative Hamiltonian systems such as Po<strong>in</strong>caré’s three-body celestial<br />

mechanics system. The other ma<strong>in</strong> classification <strong>in</strong>volves driven, damped, non-l<strong>in</strong>ear oscillatory systems.<br />

Nonl<strong>in</strong>earity and chaos is a broad and active field and thus this chapter will focus only on a few examples<br />

that illustrate the general features of non-l<strong>in</strong>ear systems. Weak non-l<strong>in</strong>earity is used to illustrate bifurcation<br />

and asymptotic attractor solutions for which the system evolves <strong>in</strong>dependent of the <strong>in</strong>itial conditions. The<br />

common s<strong>in</strong>usoidally-driven l<strong>in</strong>early-damped plane pendulum illustrates several features characteristic of the<br />

85


86 CHAPTER 4. NONLINEAR SYSTEMS AND CHAOS<br />

evolution of a non-l<strong>in</strong>ear system from order to chaos. The impact of non-l<strong>in</strong>earity on wavepacket propagation<br />

velocities and the existence of soliton solutions is discussed. The example of the three-body problem is<br />

discussed <strong>in</strong> chapter 11. The transition from lam<strong>in</strong>ar flow to turbulent flow is illustrated by fluid mechanics<br />

discussed <strong>in</strong> chapter 168. Analytic solutions of nonl<strong>in</strong>ear systems usually are not available and thus one<br />

must resort to computer simulations. As a consequence the present discussion focusses on the ma<strong>in</strong> features<br />

of the solutions for these systems and ignores how the equations of motion are solved.<br />

4.2 Weak nonl<strong>in</strong>earity<br />

Most physical oscillators become non-l<strong>in</strong>ear with <strong>in</strong>crease <strong>in</strong> amplitude of the oscillations. Consequences<br />

of non-l<strong>in</strong>earity <strong>in</strong>clude breakdown of superposition, <strong>in</strong>troduction of additional harmonics, and complicated<br />

chaotic motion that has great sensitivity to the <strong>in</strong>itial conditions as illustrated <strong>in</strong> this chapter Weak nonl<strong>in</strong>earity<br />

is <strong>in</strong>terest<strong>in</strong>g s<strong>in</strong>ce perturbation theory can be used to solve the non-l<strong>in</strong>ear equations of motion.<br />

The potential energy function for a l<strong>in</strong>ear oscillator has a pure parabolic shape about the m<strong>in</strong>imum<br />

location, that is, = 1 2 ( − 0) 2 where 0 is the location of the m<strong>in</strong>imum. Weak non-l<strong>in</strong>ear systems have<br />

small amplitude oscillations ∆ about the m<strong>in</strong>imum allow<strong>in</strong>g use of the Taylor expansion<br />

(∆) =( 0 )+∆ ( 0)<br />

<br />

By def<strong>in</strong>ition, at the m<strong>in</strong>imum (0)<br />

<br />

∆ = (∆) − ( 0 )= ∆2<br />

2!<br />

+ ∆2<br />

2!<br />

2 ( 0 )<br />

2<br />

+ ∆3<br />

3!<br />

3 ( 0 )<br />

3<br />

+ ∆4<br />

4!<br />

=0 and thus equation 41 can be written as<br />

2 ( 0 )<br />

2<br />

+ ∆3<br />

3!<br />

3 ( 0 )<br />

3<br />

+ ∆4<br />

4!<br />

4 ( 0 )<br />

4 + (4.1)<br />

4 ( 0 )<br />

4 + (4.2)<br />

2 ( 0 )<br />

For small amplitude oscillations the system is l<strong>in</strong>ear when only the second-order ∆2<br />

2! <br />

term <strong>in</strong> equation<br />

2<br />

42 is significant. The l<strong>in</strong>earity for small amplitude oscillations greatly simplifies description of the oscillatory<br />

motion <strong>in</strong> that superposition applies, and complicated chaotic motion is avoided. For slightly larger amplitude<br />

motion, where the higher-order terms <strong>in</strong> the expansion are still much smaller than the second-order term,<br />

then perturbation theory can be used as illustrated by the simple plane pendulum which is non l<strong>in</strong>ear s<strong>in</strong>ce<br />

the restor<strong>in</strong>g force equals<br />

s<strong>in</strong> ' ( − 3<br />

3! + 5<br />

5! − 7<br />

+ ) (4.3)<br />

7!<br />

This is l<strong>in</strong>ear only at very small angles where the higher-order terms <strong>in</strong> the expansion can be neglected.<br />

Consider the equation of motion at small amplitudes for the harmonically-driven, l<strong>in</strong>early-damped plane<br />

pendulum<br />

¨ + Γ˙ + <br />

2<br />

0 s<strong>in</strong> = ¨ + Γ˙ + 2 0( − 3<br />

6 )= 0 cos () (4.4)<br />

where only the first two terms <strong>in</strong> the expansion 43 have been <strong>in</strong>cluded. It was shown <strong>in</strong> chapter 3 that when<br />

s<strong>in</strong> ≈ then the steady-state solution of equation 44 is of the form<br />

() = cos ( − ) (4.5)<br />

Insert this first-order solution <strong>in</strong>to equation 44, then the cubic term <strong>in</strong> the expansion gives a term 3 =<br />

(cos 3 +3cos). Thus the perturbation expansion to third order <strong>in</strong>volves a solution of the form<br />

1<br />

4<br />

() = cos ( − )+ cos 3( − ) (4.6)<br />

This perturbation solution shows that the non-l<strong>in</strong>ear term has distorted the signal by addition of the third<br />

harmonic of the driv<strong>in</strong>g frequency with an amplitude that depends sensitively on . This illustrates that the<br />

superposition pr<strong>in</strong>ciple is not obeyed for this non-l<strong>in</strong>ear system, but, if the non-l<strong>in</strong>earity is weak, perturbation<br />

theory can be used to derive the solution of a non-l<strong>in</strong>ear equation of motion.<br />

Figure 41 illustrates that for a potential () =2 2 + 4 the 4 non-l<strong>in</strong>ear term are greatest at the<br />

maximum amplitude which makes the total energy contours <strong>in</strong> state-space more rectangular than the<br />

elliptical shape for the harmonic oscillator as shown <strong>in</strong> figure 33. The solution is of the form given <strong>in</strong><br />

equation 46.


4.2. WEAK NONLINEARITY 87<br />

Figure 4.1: The left side shows the potential energy for a symmetric potential () =2 2 + 4 . The right<br />

side shows the contours of constant total energy on a state-space diagram.<br />

4.1 Example: Non-l<strong>in</strong>ear oscillator<br />

Assume that a non-l<strong>in</strong>ear oscillator has a potential given by<br />

() = 2<br />

2 − 3<br />

3<br />

where is small. F<strong>in</strong>d the solution of the equation of motion to first order <strong>in</strong> , assum<strong>in</strong>g =0at =0.<br />

The equation of motion for the nonl<strong>in</strong>ear oscillator is<br />

¨ = − = − + 2<br />

<br />

If the 2 term is neglected, then the second-order equation of motion reduces to a normal l<strong>in</strong>ear oscillator<br />

with<br />

0 = s<strong>in</strong> ( 0 + )<br />

where<br />

r<br />

<br />

0 =<br />

<br />

Assume that the first-order solution has the form<br />

1 = 0 + 1<br />

Substitut<strong>in</strong>g this <strong>in</strong>to the equation of motion, and neglect<strong>in</strong>g terms of higher order than gives<br />

¨ 1 + 2 0 1 = 2 0 = 2<br />

2 [1 − cos (2 0)]<br />

To solve this try a particular <strong>in</strong>tegral<br />

1 = + cos (2 0 )<br />

and substitute <strong>in</strong>to the equation of motion gives<br />

−3 2 0 cos (2 0 )+ 2 0 = 2<br />

2 − 2<br />

2 cos (2 0)<br />

Comparison of the coefficients gives<br />

= 2<br />

2 2 0<br />

= 2<br />

6 2 0


88 CHAPTER 4. NONLINEAR SYSTEMS AND CHAOS<br />

The homogeneous equation is<br />

¨ 1 + 2 0 1 =0<br />

which has a solution of the form<br />

1 = 1 s<strong>in</strong> ( 0 )+ 2 cos ( 0 )<br />

Thus comb<strong>in</strong><strong>in</strong>g the particular and homogeneous solutions gives<br />

∙ ¸<br />

<br />

2<br />

1 =( + 1 )s<strong>in</strong>( 0 )+<br />

2 2 + 2 cos ( 0 )+ 2<br />

0<br />

6 2 cos (2 0 )<br />

0<br />

The <strong>in</strong>itial condition =0at =0then gives<br />

and<br />

1 =( + 1 )s<strong>in</strong>( 0 )+ 2<br />

2 0<br />

2 = − 22<br />

3 2<br />

∙ 1<br />

2 − 2 3 cos ( 0)+ 1 6 cos (2 0)¸<br />

The constant ( + 1 ) is given by the <strong>in</strong>itial amplitude and velocity.<br />

This system is nonl<strong>in</strong>ear <strong>in</strong> that the output amplitude is not proportional to the <strong>in</strong>put amplitude. <strong>Second</strong>ly,<br />

a large amplitude second harmonic component is <strong>in</strong>troduced <strong>in</strong> the output waveform; that is, for a non-l<strong>in</strong>ear<br />

system the ga<strong>in</strong> and frequency decomposition of the output differs from the <strong>in</strong>put. Note that the frequency<br />

composition is amplitude dependent. This particular example of a nonl<strong>in</strong>ear system does not exhibit chaos.<br />

The Laboratory for Laser Energetics uses nonl<strong>in</strong>ear crystals to double the frequency of laser light.<br />

4.3 Bifurcation, and po<strong>in</strong>t attractors<br />

Interest<strong>in</strong>g new phenomena, such as bifurcation, and attractors, occur when the non-l<strong>in</strong>earity is large. In<br />

chapter 3 it was shown that the state-space diagram ( ˙ ) for an undamped harmonic oscillator is an<br />

ellipse with dimensions def<strong>in</strong>ed by the total energy of the system. As shown <strong>in</strong> figure 35 for the damped<br />

harmonic oscillator, the state-space diagram spirals <strong>in</strong>wards to the orig<strong>in</strong> due to dissipation of energy. Nonl<strong>in</strong>earity<br />

distorts the shape of the ellipse or spiral on the state-space diagram, and thus the state-space, or<br />

correspond<strong>in</strong>g phase-space, diagrams, provide useful representations of the motion of l<strong>in</strong>ear and non-l<strong>in</strong>ear<br />

periodic systems.<br />

The complicated motion of non-l<strong>in</strong>ear systems makes it necessary to dist<strong>in</strong>guish between transient and<br />

asymptotic behavior. The damped harmonic oscillator executes a transient spiral motion that asymptotically<br />

approaches the orig<strong>in</strong>. The transient behavior depends on the <strong>in</strong>itial conditions, whereas the asymptotic limit<br />

of the steady-state solution is a specific location, that is called a po<strong>in</strong>t attractor. The po<strong>in</strong>t attractor for<br />

damped motion <strong>in</strong> the anharmonic potential well<br />

() =2 2 + 4 (4.7)<br />

is at the m<strong>in</strong>imum, which is the orig<strong>in</strong> of the state-space diagram as shown <strong>in</strong> figure 41.<br />

The more complicated one-dimensional potential well<br />

() =8− 4 2 +05 4 (4.8)<br />

shown <strong>in</strong> figure 42 has two m<strong>in</strong>ima that are symmetric about =0with a saddle of height 8.<br />

The k<strong>in</strong>etic plus potential energies of a particle with mass =2 released <strong>in</strong> this potential, will be<br />

assumed to be given by<br />

( ˙) = ˙ 2 + () (4.9)<br />

The state-space plot <strong>in</strong> figure 42 shows contours of constant energy with the m<strong>in</strong>ima at ( ˙) =(±2 0).<br />

At slightly higher total energy the contours are closed loops around either of the two m<strong>in</strong>ima at = ±2.<br />

At total energies above the saddle energy of 8 the contours are peanut-shaped and are symmetric about<br />

the orig<strong>in</strong>. Assum<strong>in</strong>g that the motion is weakly damped, then a particle released with total energy <br />

which is higher than will follow a peanut-shaped spiral trajectory centered at ( ˙) =(0 0) <strong>in</strong> the


4.4. LIMIT CYCLES 89<br />

Figure 4.2: The left side shows the potential energy for a bimodal symmetric potential () =8− 4 2 +<br />

05 4 . The right-hand figure shows contours of the sum of k<strong>in</strong>etic and potential energies on a state-space<br />

diagram. For total energies above the saddle po<strong>in</strong>t the particle follows peanut-shaped trajectories <strong>in</strong> statespace<br />

centered around ( ˙) =(0 0). For total energies below the saddle po<strong>in</strong>t the particle will have closed<br />

trajectories about either of the two symmetric m<strong>in</strong>ima located at ( ˙) =(±2 0). Thus the system solution<br />

bifurcates when the total energy is below the saddle po<strong>in</strong>t.<br />

state-space diagram for . For there are two separate solutions for the two<br />

m<strong>in</strong>imum centered at = ±2 and ˙ =0. Thisisanexampleofbifurcation where the one solution for<br />

bifurcates <strong>in</strong>to either of the two solutions for .<br />

For an <strong>in</strong>itial total energy damp<strong>in</strong>g will result <strong>in</strong> spiral trajectories of the particle that<br />

will be trapped <strong>in</strong> one of the two m<strong>in</strong>ima. For the particle trajectories are centered giv<strong>in</strong>g<br />

the impression that they will term<strong>in</strong>ate at ( ˙) =(0 0) when the k<strong>in</strong>etic energy is dissipated. However, for<br />

the particle will be trapped <strong>in</strong> one of the two m<strong>in</strong>imum and the trajectory will term<strong>in</strong>ate<br />

at the bottom of that potential energy m<strong>in</strong>imum occurr<strong>in</strong>g at ( ˙) =(±2 0). These two possible term<strong>in</strong>al<br />

po<strong>in</strong>ts of the trajectory are called po<strong>in</strong>t attractors. This example appears to have a s<strong>in</strong>gle attractor for<br />

which bifurcates lead<strong>in</strong>g to two attractors at ( ˙) =(±2 0) for . The<br />

determ<strong>in</strong>ation as to which m<strong>in</strong>imum traps a given particle depends on exactly where the particle starts <strong>in</strong><br />

state space and the damp<strong>in</strong>g etc. That is, for this case, where there is symmetry about the -axis, the<br />

particle has an <strong>in</strong>itial total energy then the <strong>in</strong>itial conditions with radians of state space<br />

will lead to trajectories that are trapped <strong>in</strong> the left m<strong>in</strong>imum, and the other radians of state space will be<br />

trapped <strong>in</strong> the right m<strong>in</strong>imum. Trajectories start<strong>in</strong>g near the split between these two halves of the start<strong>in</strong>g<br />

state space will be sensitive to the exact start<strong>in</strong>g phase. This is an example of sensitivity to <strong>in</strong>itial conditions.<br />

4.4 Limit cycles<br />

4.4.1 Po<strong>in</strong>caré-Bendixson theorem<br />

Coupled first-order differential equations <strong>in</strong> two dimensions of the form<br />

˙ = ( ) ˙ = ( ) (4.10)<br />

occur frequently <strong>in</strong> physics. The state-space paths do not cross for such two-dimensional autonomous systems,<br />

where an autonomous system is not explicitly dependent on time.<br />

The Po<strong>in</strong>caré-Bendixson theorem states that, state-space, and phase-space, can have three possible paths:<br />

(1) closed paths, like the elliptical paths for the undamped harmonic oscillator,<br />

(2) term<strong>in</strong>ate at an equilibrium po<strong>in</strong>t as →∞, like the po<strong>in</strong>t attractor for a damped harmonic oscillator,<br />

(3) tend to a limit cycle as →∞.<br />

The limit cycle is unusual <strong>in</strong> that the periodic motion tends asymptotically to the limit-cycle attractor<br />

<strong>in</strong>dependent of whether the <strong>in</strong>itial values are <strong>in</strong>side or outside the limit cycle. The balance of dissipative forces<br />

and driv<strong>in</strong>g forces often leads to limit-cycle attractors, especially <strong>in</strong> biological applications. Identification of<br />

limit-cycle attractors, as well as the trajectories of the motion towards these limit-cycle attractors, is more<br />

complicated than for po<strong>in</strong>t attractors.


90 CHAPTER 4. NONLINEAR SYSTEMS AND CHAOS<br />

Figure 4.3: The Po<strong>in</strong>caré-Bendixson theorem allows the follow<strong>in</strong>g three scenarios for two-dimensional autonomous<br />

systems. (1) Closed paths as illustrated by the undamped harmonic oscillator. (2) Term<strong>in</strong>ate at<br />

an equilibrium po<strong>in</strong>t as →∞, as illustrated by the damped harmonic oscillator, and (3) Tend to a limit<br />

cycle as →∞as illustrated by the van der Pol oscillator.<br />

4.4.2 van der Pol damped harmonic oscillator:<br />

The van der Pol damped harmonic oscillator illustrates a non-l<strong>in</strong>ear equation that leads to a well-studied,<br />

limit-cycle attractor that has important applications <strong>in</strong> diverse fields. The van der Pol oscillator has an<br />

equation of motion given by<br />

2 <br />

2 + ¡ 2 − 1 ¢ <br />

+ 2 0 =0 (4.11)<br />

The non-l<strong>in</strong>ear ¡ 2 − 1 ¢ <br />

<br />

damp<strong>in</strong>g term is unusual <strong>in</strong> that the sign changes when =1lead<strong>in</strong>g to<br />

positive damp<strong>in</strong>g for 1 and negative damp<strong>in</strong>g for 1 To simplify equation 411 assume that the term<br />

2 0 = that is, 2 0 =1.<br />

This equation was studied extensively dur<strong>in</strong>g the 1920’s and 1930’s by the Dutch eng<strong>in</strong>eer, Balthazar<br />

van der Pol, for describ<strong>in</strong>g electronic circuits that <strong>in</strong>corporate feedback. The form of the solution can be<br />

simplified by def<strong>in</strong><strong>in</strong>g a variable ≡ <br />

<br />

Then the second-order equation 411 can be expressed as two<br />

coupled first-order equations.<br />

≡ <br />

(4.12)<br />

<br />

<br />

= − − ¡ 2 − 1 ¢ (4.13)<br />

<br />

It is advantageous to transform the ( ˙ ) state space to polar coord<strong>in</strong>ates by sett<strong>in</strong>g<br />

and us<strong>in</strong>g the fact that 2 = 2 + 2 Therefore<br />

= cos (4.14)<br />

= s<strong>in</strong> <br />

<br />

= + <br />

<br />

Similarly for the angle coord<strong>in</strong>ate<br />

<br />

<br />

<br />

<br />

Multiply equation 416 by and 417 by and subtract gives<br />

(4.15)<br />

= <br />

cos − s<strong>in</strong> (4.16)<br />

<br />

= <br />

s<strong>in</strong> + cos (4.17)<br />

<br />

2 <br />

= − <br />

<br />

(4.18)


4.4. LIMIT CYCLES 91<br />

Figure 4.4: Solutions of the van der Pol system for =02 top row and =5bottom row, assum<strong>in</strong>g that<br />

2 0 =1. The left column shows the time dependence (). The right column shows the correspond<strong>in</strong>g ( ˙)<br />

state space plots. Upper: Weak nonl<strong>in</strong>earity, = 02; At large times the solution tends to one limit<br />

cycle for <strong>in</strong>itial values <strong>in</strong>side or outside the limit cycle attractor. The amplitude () for two <strong>in</strong>itial conditions<br />

approaches an approximately harmonic oscillation. Lower: Strong nonl<strong>in</strong>earity, μ = 5; Solutions<br />

approach a common limit cycle attractor for <strong>in</strong>itial values <strong>in</strong>side or outside the limit cycle attractor while<br />

the amplitude () approaches a common approximate square-wave oscillation.<br />

Equations 415 and 418 allow the van der Pol equations of motion to be written <strong>in</strong> polar coord<strong>in</strong>ates<br />

<br />

<br />

= − ¡ 2 cos 2 − 1 ¢ s<strong>in</strong> 2 (4.19)<br />

<br />

<br />

= −1 − ¡ 2 cos 2 − 1 ¢ s<strong>in</strong> cos (4.20)<br />

The non-l<strong>in</strong>ear terms on the right-hand side of equations 419 − 20 have a complicated form.<br />

Weak non-l<strong>in</strong>earity: 1<br />

In the limit that → 0, equations419 420 correspond to a circular state-space trajectory similar to the<br />

harmonic oscillator. That is, the solution is of the form<br />

() = s<strong>in</strong> ( − 0 ) (4.21)<br />

where and 0 are arbitrary parameters. For weak non-l<strong>in</strong>earity, 1 the angular equation 420 has a<br />

rotational frequency that is unity s<strong>in</strong>ce the s<strong>in</strong> cos term changes sign twice per period, <strong>in</strong> addition to the


92 CHAPTER 4. NONLINEAR SYSTEMS AND CHAOS<br />

small value of . For1 and 1 the radial equation 419 has a sign of the ¡ 2 cos 2 − 1 ¢ term that<br />

is positive and thus the radius <strong>in</strong>creases monotonically to unity. For 1 the bracket is predom<strong>in</strong>antly<br />

negative result<strong>in</strong>g <strong>in</strong> a spiral decrease <strong>in</strong> the radius. Thus, for very weak non-l<strong>in</strong>earity, this radial behavior<br />

results <strong>in</strong> the amplitude spirall<strong>in</strong>g to a well def<strong>in</strong>ed limit-cycle attractor value of =2as illustrated by<br />

the state-space plots <strong>in</strong> figure 44 for cases where the <strong>in</strong>itial condition is <strong>in</strong>side or external to the circular<br />

attractor. The f<strong>in</strong>al amplitude for different <strong>in</strong>itial conditions also approach the same asymptotic behavior.<br />

Dom<strong>in</strong>ant non-l<strong>in</strong>earity: 1<br />

For the case where the non-l<strong>in</strong>earity is dom<strong>in</strong>ant, that is 1, then as shown <strong>in</strong> figure 44, the system<br />

approaches a well def<strong>in</strong>ed attractor, but <strong>in</strong> this case it has a significantly skewed shape <strong>in</strong> state-space, while<br />

the amplitude approximates a square wave. The solution rema<strong>in</strong>s close to =+2until = ˙ ≈ +7 and<br />

then it relaxes quickly to = −2 with = ˙ ≈ 0 This is followed by the mirror image. This behavior is<br />

called a relaxed vibration <strong>in</strong> that a tension builds up slowly then dissipates by a sudden relaxation process.<br />

The seesaw is an extreme example of a relaxation oscillator where the seesaw angle switches spontaneously<br />

from one solution to the other when the difference <strong>in</strong> their moment arms changes sign.<br />

The study of feedback <strong>in</strong> electronic circuits was the stimulus for study of this equation by van der<br />

Pol. However, Lord Rayleigh first identified such relaxation oscillator behavior <strong>in</strong> 1880 dur<strong>in</strong>g studies of<br />

vibrations of a str<strong>in</strong>ged <strong>in</strong>strument excited by a bow, or the squeak<strong>in</strong>g of a brake drum. In his discussion of<br />

non-l<strong>in</strong>ear effects <strong>in</strong> acoustics, he derived the equation<br />

¨ − ( − ˙ 2 ) ˙ + 2 0 (4.22)<br />

Differentiation of Rayleigh’s equation 422 gives<br />

...<br />

− ( − 3 ˙ 2 )¨ + 2 0 ˙ =0 (4.23)<br />

Us<strong>in</strong>g the substitution of<br />

leads to the relations<br />

= 0<br />

r<br />

3<br />

<br />

˙ (4.24)<br />

r r r <br />

˙<br />

... ¨<br />

˙ =<br />

¨ = =<br />

(4.25)<br />

3 0 3 0 3 0<br />

Substitut<strong>in</strong>g these relations <strong>in</strong>to equation 423 gives<br />

r r ∙<br />

¨ <br />

− − 3 ˙ 2 ¸ r ˙ <br />

+ 2 <br />

0 =0 (4.26)<br />

3 0 3 0 3 0<br />

Multiply<strong>in</strong>g by 0<br />

q<br />

3<br />

<br />

and rearrang<strong>in</strong>g leads to the van der Pol equation<br />

2 0<br />

¨ − 0<br />

2 (0 2 − 2 ) ˙ − 2 0 =0 (4.27)<br />

The rhythm of a heartbeat driven by a pacemaker is an important application where the self-stabilization of<br />

the attractor is a desirable characteristic to stabilize an irregular heartbeat; the medical term is arrhythmia.<br />

The mechanism that leads to synchronization of the many pacemaker cells <strong>in</strong> the heart and human body due<br />

to the <strong>in</strong>fluence of an implanted pacemaker is discussed <strong>in</strong> chapter 1412. Another biological application of<br />

limit cycles is the time variation of animal populations.<br />

In summary the non-l<strong>in</strong>ear damp<strong>in</strong>g of the van der Pol oscillator leads to a self-stabilized, s<strong>in</strong>gle limitcycle<br />

attractor that is <strong>in</strong>sensitive to the <strong>in</strong>itial conditions. The van der Pol oscillator has many important<br />

applications such as bowed musical <strong>in</strong>struments, electrical circuits, and human anatomy as mentioned above.<br />

The van der Pol oscillator illustrates the complicated manifestations of the motion that can be exhibited by<br />

non-l<strong>in</strong>ear systems


4.5. HARMONICALLY-DRIVEN, LINEARLY-DAMPED, PLANE PENDULUM 93<br />

4.5 Harmonically-driven, l<strong>in</strong>early-damped, plane pendulum<br />

The harmonically-driven, l<strong>in</strong>early-damped, plane pendulum illustrates many of the phenomena exhibited by<br />

non-l<strong>in</strong>ear systems as they evolve from ordered to chaotic motion. It illustrates the remarkable fact that<br />

determ<strong>in</strong>ism does not imply either regular behavior or predictability. The well-known, harmonically-driven<br />

l<strong>in</strong>early-damped pendulum provides an ideal basis for an <strong>in</strong>troduction to non-l<strong>in</strong>ear dynamics 1 .<br />

Consider a harmonically-driven l<strong>in</strong>early-damped plane pendulum of moment of <strong>in</strong>ertia and mass <strong>in</strong><br />

a gravitational field that is driven by a torque due to a force () = cos act<strong>in</strong>g at a moment arm .<br />

The damp<strong>in</strong>g term is and the angular displacement of the pendulum, relative to the vertical, is . The<br />

equation of motion of the harmonically-driven l<strong>in</strong>early-damped simple pendulum can be written as<br />

¨ + ˙ + s<strong>in</strong> = cos (4.28)<br />

Note that the s<strong>in</strong>usoidal restor<strong>in</strong>g force for the plane pendulum is non-l<strong>in</strong>ear for large angles . Thenatural<br />

period of the free pendulum is<br />

r<br />

<br />

0 =<br />

(4.29)<br />

<br />

A dimensionless parameter , which is called the drive strength, is def<strong>in</strong>ed by<br />

≡ <br />

(4.30)<br />

<br />

The equation of motion 428 can be generalized by <strong>in</strong>troduc<strong>in</strong>g dimensionless units for both time ˜ and<br />

relative drive frequency ˜ def<strong>in</strong>ed by<br />

˜ ≡ 0 ˜ ≡ 0<br />

(4.31)<br />

In addition, def<strong>in</strong>e the <strong>in</strong>verse damp<strong>in</strong>g factor as<br />

≡ 0<br />

(4.32)<br />

<br />

These def<strong>in</strong>itions allow equation 428 to be written <strong>in</strong> the dimensionless form<br />

2 <br />

˜ + 1 <br />

2 ˜ +s<strong>in</strong> = cos ˜˜ (4.33)<br />

The behavior of the angle for the driven damped plane pendulum depends on the drive strength <br />

and the damp<strong>in</strong>g factor . Consider the case where equation 433 is evaluated assum<strong>in</strong>g that the damp<strong>in</strong>g<br />

coefficient =2, and that the relative angular frequency ˜ = 2 3 which is close to resonance where chaotic<br />

phenomena are manifest. The Runge-Kutta method is used to solve this non-l<strong>in</strong>ear equation of motion.<br />

4.5.1 Close to l<strong>in</strong>earity<br />

For drive strength =02 the amplitude is sufficiently small that s<strong>in</strong> ' superposition applies, and the<br />

solution is identical to that for the driven l<strong>in</strong>early-damped l<strong>in</strong>ear oscillator. As shown <strong>in</strong> figure 45, once<br />

the transient solution dies away, the steady-state solution asymptotically approaches one attractor that has<br />

an amplitude of ±03 radians and a phase shift with respect to the driv<strong>in</strong>g force. The abscissa is given<br />

<strong>in</strong> units of the dimensionless time ˜ = 0 . The transient solution depends on the <strong>in</strong>itial conditions and<br />

dies away after about 5 periods, whereas the steady-state solution is <strong>in</strong>dependent of the <strong>in</strong>itial conditions<br />

and has a state-space diagram that has an elliptical shape, characteristic of the harmonic oscillator. For all<br />

<strong>in</strong>itial conditions, the time dependence and state space diagram for steady-state motion approaches a unique<br />

solution,calledan“attractor”, that is, the pendulum oscillates s<strong>in</strong>usoidally with a given amplitude at the<br />

frequency of the driv<strong>in</strong>g force and with a constant phase shift , i.e.<br />

() = cos( − ) (4.34)<br />

This solution is identical to that for the harmonically-driven, l<strong>in</strong>early-damped, l<strong>in</strong>ear oscillator discussed <strong>in</strong><br />

chapter 36<br />

1 A similar approach is used by the book "Chaotic Dynamics" by Baker and Gollub[Bak96].


94 CHAPTER 4. NONLINEAR SYSTEMS AND CHAOS<br />

Figure 4.5: Motion of the driven damped pendulum for drive strengths of =02, =09 =105 and<br />

=1078. The left side shows the time dependence of the deflection angle with the time axis expressed<br />

<strong>in</strong> dimensionless units ˜. The right side shows the correspond<strong>in</strong>g state-space plots. These plots assume<br />

˜ = 0<br />

= 2 3<br />

, =2, and the motion starts with = =0.


4.5. HARMONICALLY-DRIVEN, LINEARLY-DAMPED, PLANE PENDULUM 95<br />

20<br />

3<br />

2<br />

10<br />

1<br />

2 4 6 8 10 12 14<br />

t<br />

2 2<br />

1<br />

2<br />

10<br />

20<br />

3<br />

2<br />

10<br />

1<br />

2 4 6 8 10 12 14<br />

t<br />

2 2<br />

1<br />

2<br />

10<br />

20<br />

Figure 4.6: The driven damped pendulum assum<strong>in</strong>g that ˜ = 2 3 , =2, with <strong>in</strong>itial conditions (0) = − 2 ,<br />

(0) = 0. The system exhibits period-two motion for drive strengths of =1078 as shown by the state<br />

space diagram for cycles 10 − 20. For =1081 the system exhibits period-four motion shown for cycles<br />

10 − 30.<br />

4.5.2 Weak nonl<strong>in</strong>earity<br />

Figure 45 shows that for drive strength =09, after the transient solution dies away, the steady-state<br />

solution settles down to one attractor that oscillates at the drive frequency with an amplitude of slightly<br />

more than 2<br />

radians for which the small angle approximation fails. The distortion due to the non-l<strong>in</strong>earity<br />

is exhibited by the non-elliptical shape of the state-space diagram.<br />

The observed behavior can be calculated us<strong>in</strong>g the successive approximation method discussed <strong>in</strong> chapter<br />

42. That is, close to small angles the s<strong>in</strong>e function can be approximated by replac<strong>in</strong>g<br />

s<strong>in</strong> ≈ − 1 6 3<br />

<strong>in</strong> equation 433 to give<br />

1<br />

¨ +<br />

˙ + 2 0<br />

µ − 1 <br />

6 3 = cos ˜˜ (4.35)<br />

As a first approximation assume that<br />

(˜) ≈ cos(˜˜ − )<br />

then the small 3 term <strong>in</strong> equation 435 contributes a term proportional to cos 3 (˜˜ − ). But<br />

cos 3 (˜˜ − ) = 1 ¡<br />

cos 3(˜˜ − )+3cos(˜˜ − ) ¢<br />

4<br />

That is, the nonl<strong>in</strong>earity <strong>in</strong>troduces a small term proportional to cos 3( − ). S<strong>in</strong>ce the right-hand side of<br />

equation 435 is a function of only cos then the terms <strong>in</strong> ˙ and ¨ on the left hand side must conta<strong>in</strong><br />

the third harmonic cos 3( − ) term. Thus a better approximation to the solution is of the form<br />

(˜) = £ cos(˜˜ − )+ cos 3(˜˜ − ) ¤ (4.36)


96 CHAPTER 4. NONLINEAR SYSTEMS AND CHAOS<br />

where the admixture coefficient 1. This successive approximation method can be repeated to add<br />

additional terms proportional to cos ( − ) where is an <strong>in</strong>teger with ≥ 3. Thus the nonl<strong>in</strong>earity<br />

<strong>in</strong>troduces progressively weaker -fold harmonics to the solution. This successive approximation approach<br />

is viable only when the admixture coefficient 1 Note that these harmonics are <strong>in</strong>teger multiples of ,<br />

thus the steady-state response is identical for each full period even though the state space contours deviate<br />

from an elliptical shape.<br />

4.5.3 Onset of complication<br />

Figure 45 shows that for =105 the drive strength is sufficiently strong to cause the transient solution for<br />

the pendulum to rotate through two complete cycles before settl<strong>in</strong>g down to a s<strong>in</strong>gle steady-state attractor<br />

solution at the drive frequency. However, this attractor solution is shifted two complete rotations relative<br />

to the <strong>in</strong>itial condition. The state space diagram clearly shows the roll<strong>in</strong>g motion of the transient solution<br />

for the first two periods prior to the system settl<strong>in</strong>g down to a s<strong>in</strong>gle steady-state attractor. The successive<br />

approximation approach completely fails at this coupl<strong>in</strong>g strength s<strong>in</strong>ce oscillates through large values that<br />

are multiples of <br />

Figure 45 shows that for drive strength =1078 the motion evolves to a much more complicated<br />

periodic motion with a period that is three times the period of the driv<strong>in</strong>g force. Moreover the amplitude<br />

exceeds 2 correspond<strong>in</strong>g to the pendulum oscillat<strong>in</strong>g over top dead center with the centroid of the motion<br />

offset by 3 from the <strong>in</strong>itial condition. Both the state-space diagram, and the time dependence of the motion,<br />

illustrate the complexity of this motion which depends sensitively on the magnitude of the drive strength <br />

<strong>in</strong> addition to the <strong>in</strong>itial conditions, ((0)(0)) and damp<strong>in</strong>g factor as is shown <strong>in</strong> figure 46<br />

4.5.4 Period doubl<strong>in</strong>g and bifurcation<br />

For drive strength =1078 with the <strong>in</strong>itial condition ((0)(0)) = (0 0) the system exhibits a regular<br />

motion with a period that is three times the drive period. In contrast, if the <strong>in</strong>itial condition is [(0) =<br />

− 2<br />

(0) = 0] then, as shown <strong>in</strong> figure 46 the steady-state solution has the drive frequency with no offset<br />

<strong>in</strong> , that is, it exhibits period-one oscillation. This appearance of two separate and very different attractors<br />

for =1078 us<strong>in</strong>g different <strong>in</strong>itial conditions, is called bifurcation.<br />

An additional feature of the system response for =1078 is that chang<strong>in</strong>g the <strong>in</strong>itial conditions to<br />

[(0) = − 2<br />

(0) = 0] shows that the amplitude of the even and odd periods of oscillation differ slightly<br />

<strong>in</strong> shape and amplitude, that is, the system really has period-two oscillation. This period-two motion, i.e.<br />

period doubl<strong>in</strong>g, is clearly illustrated by the state space diagram <strong>in</strong> that, although the motion still is<br />

dom<strong>in</strong>ated by period-one oscillations, the even and odd cycles are slightly displaced. Thus, for different<br />

<strong>in</strong>itial conditions, the system for =1078 bifurcates <strong>in</strong>to either of two attractors that have very different<br />

waveforms, one of which exhibits period doubl<strong>in</strong>g.<br />

The period doubl<strong>in</strong>g exhibited for =1078 is followed by a second period doubl<strong>in</strong>g when =1081 as<br />

shown <strong>in</strong> figure 46 . With <strong>in</strong>crease <strong>in</strong> drive strength this period doubl<strong>in</strong>g keeps <strong>in</strong>creas<strong>in</strong>g <strong>in</strong> b<strong>in</strong>ary multiples<br />

to period 8, 16, 32, 64 etc. Numerically it is found that the threshold for period doubl<strong>in</strong>g is 1 =10663<br />

fromtwotofouroccursat 2 =10793 etc. Feigenbaum showed that this cascade <strong>in</strong>creases with <strong>in</strong>crease <strong>in</strong><br />

drive strength accord<strong>in</strong>g to the relation that obeys<br />

( +1 − ) ' 1 ( − −1 ) (4.37)<br />

where =46692016, is called a Feigenbaum number. As →∞this cascad<strong>in</strong>g sequence goes to a limit<br />

where<br />

=10829 (4.38)<br />

4.5.5 Roll<strong>in</strong>g motion<br />

It was shown that for 105 the transient solution causes the pendulum to have angle excursions exceed<strong>in</strong>g<br />

2, that is, the system rolls over top dead center. For drive strengths <strong>in</strong> the range 13 14 the steadystate<br />

solution for the system undergoes cont<strong>in</strong>uous roll<strong>in</strong>g motion as illustrated <strong>in</strong> figure 47. The time<br />

dependence for the angle exhibits a periodic oscillatory motion superimposed upon a monotonic roll<strong>in</strong>g<br />

motion, whereas the time dependence of the angular frequency = <br />

<br />

is periodic. The state space plots


4.5. HARMONICALLY-DRIVEN, LINEARLY-DAMPED, PLANE PENDULUM 97<br />

Figure 4.7: Roll<strong>in</strong>g motion for the driven damped plane pendulum for =14. (a) The time dependence<br />

of angle () <strong>in</strong>creases by 2 per drive period whereas (b) the angular velocity () exhibits periodicity. (c)<br />

The state space plot for roll<strong>in</strong>g motion is shown with the orig<strong>in</strong> shifted by 2 per revolution to keep the plot<br />

with<strong>in</strong> the bounds − +<br />

for roll<strong>in</strong>g motion corresponds to a cha<strong>in</strong> of loops with a spac<strong>in</strong>g of 2 between each loop. The state space<br />

diagram for roll<strong>in</strong>g motion is more compactly presented if the orig<strong>in</strong> is shifted by 2 per revolution to keep<br />

the plot with<strong>in</strong> bounds as illustrated <strong>in</strong> figure 47.<br />

4.5.6 Onset of chaos<br />

When the drive strength is <strong>in</strong>creased to =1105 then the system does not approach a unique attractor<br />

as illustrated by figure 48 which shows state space orbits for cycles 25 − 200. Note that these orbits do<br />

not repeat imply<strong>in</strong>g the onset of chaos. For drive strengths greater than =10829 the driven damped<br />

plane pendulum starts to exhibit chaotic behavior. The onset of chaotic motion is illustrated by mak<strong>in</strong>g a 3-<br />

dimensional plot which comb<strong>in</strong>es the time coord<strong>in</strong>ate with the state-space coord<strong>in</strong>ates as illustrated <strong>in</strong> figure<br />

48. This plot shows 16 trajectories start<strong>in</strong>g at different <strong>in</strong>itial values <strong>in</strong> the range −015 015<br />

for =1168. Some solutions are erratic <strong>in</strong> that, while try<strong>in</strong>g to oscillate at the drive frequency, they never<br />

settle down to a steady periodic motion which is characteristic of chaotic motion. Figure 48 illustrates<br />

the considerable sensitivity of the motion to the <strong>in</strong>itial conditions. That is, this determ<strong>in</strong>istic system can<br />

exhibit either order, or chaos, dependent on m<strong>in</strong>iscule differences <strong>in</strong> <strong>in</strong>itial conditions.<br />

Figure 4.8: Left: Space-space orbits for the driven damped pendulum with =1105. Note that the orbits<br />

do not repeat for cycles 25 to 200. Right: Time-state-space diagram for =1168. The plot shows 16<br />

trajectories start<strong>in</strong>g with different <strong>in</strong>itial values <strong>in</strong> the range −015 015.


98 CHAPTER 4. NONLINEAR SYSTEMS AND CHAOS<br />

Figure 4.9: State-space plots for the harmonically-driven, l<strong>in</strong>early-damped, pendulum for driv<strong>in</strong>g amplitudes<br />

of =05 and =12. These calculations were performed us<strong>in</strong>g the Runge-Kutta method by E. Shah,<br />

(Private communication)<br />

4.6 Differentiation between ordered and chaotic motion<br />

Chapter 45 showed that motion <strong>in</strong> non-l<strong>in</strong>ear systems can exhibit both order and chaos. The transition<br />

between ordered motion and chaotic motion depends sensitively on both the <strong>in</strong>itial conditions and the model<br />

parameters. It is surpris<strong>in</strong>gly difficult to unambiguously dist<strong>in</strong>guish between complicated ordered motion<br />

and chaotic motion. Moreover, the motion can fluctuate between order and chaos <strong>in</strong> an erratic manner<br />

depend<strong>in</strong>g on the <strong>in</strong>itial conditions. The extremely sensitivity to <strong>in</strong>itial conditions of the motion for nonl<strong>in</strong>ear<br />

systems, makes it essential to have quantitative measures that can characterize the degree of order, and<br />

<strong>in</strong>terpret the complicated dynamical motion of systems. As an illustration, consider the harmonically-driven,<br />

l<strong>in</strong>early-damped, pendulum with =2 and driv<strong>in</strong>g force () = s<strong>in</strong> ˜˜ where ˜ = 2 3<br />

.Figure49 shows<br />

the state-space plots for two driv<strong>in</strong>g amplitudes, =05 which leads to ordered motion, and =12<br />

which leads to possible chaotic motion. It can be seen that for =05 the state-space diagram converges<br />

to a s<strong>in</strong>gle attractor once the transient solution has died away. This is <strong>in</strong> contrast to the case for =12<br />

where the state-space diagram does not converge to a s<strong>in</strong>gle attractor, but exhibits possible chaotic motion.<br />

Three quantitative measures can be used to differentiate ordered motion from chaotic motion for this system;<br />

namely, the Lyapunov exponent, the bifurcation diagram, and the Po<strong>in</strong>caré section, as illustrated below.<br />

4.6.1 Lyapunov exponent<br />

The Lyapunov exponent provides a quantitative and useful measure of the <strong>in</strong>stability of trajectories, and how<br />

quickly nearby <strong>in</strong>itial conditions diverge. It compares two identical systems that start with an <strong>in</strong>f<strong>in</strong>itesimally<br />

small difference <strong>in</strong> the <strong>in</strong>itial conditions <strong>in</strong> order to ascerta<strong>in</strong> whether they converge to the same attractor<br />

at long times, correspond<strong>in</strong>g to a stable system, or whether they diverge to very different attractors, characteristic<br />

of chaotic motion. If the <strong>in</strong>itial separation between the trajectories <strong>in</strong> phase space at =0is | 0 |,<br />

then to first order the time dependence of the difference can be assumed to depend exponentially on time.<br />

That is,<br />

|()| ∼ | 0 | (4.39)<br />

where is the Lyapunov exponent. That is, the Lyapunov exponent is def<strong>in</strong>ed to be<br />

= lim<br />

lim<br />

→∞ 0 →0<br />

1 |()|<br />

ln<br />

| 0 |<br />

(4.40)<br />

Systems for which the Lyapunov exponent 0 (negative), converge exponentially to the same attractor<br />

solution at long times s<strong>in</strong>ce |()| → 0 for →∞. By contrast, systems for which 0 (positive) diverge<br />

to completely different long-time solutions, that is, |()| → ∞ for → ∞. Even for <strong>in</strong>f<strong>in</strong>itesimally


4.6. DIFFERENTIATION BETWEEN ORDERED AND CHAOTIC MOTION 99<br />

Figure 4.10: Lyapunov plots of ∆ versus time for two <strong>in</strong>itial start<strong>in</strong>g po<strong>in</strong>ts differ<strong>in</strong>g by ∆ 0 =0001.<br />

The parameters are =2 and () = s<strong>in</strong>( 2 3 ) and ∆ =004. The Lyapunov exponent for =05<br />

which is drawn as a dashed l<strong>in</strong>e, is convergent with = −0251 For =12 the exponent is divergent as<br />

<strong>in</strong>dicated by the dashed l<strong>in</strong>e which as a slope of =01538 These calculations were performed us<strong>in</strong>g the<br />

Runge-Kutta method by E. Shah, (Private communication)<br />

small differences <strong>in</strong> the <strong>in</strong>itial conditions, systems hav<strong>in</strong>g a positive Lyapunov exponent diverge to different<br />

attractors, whereas when the Lyapunov exponent 0 they correspond to stable solutions.<br />

Figure 410 illustrates Lyapunov plots for the harmonically-driven, l<strong>in</strong>early-damped, plane pendulum,<br />

with the same conditions discussed <strong>in</strong> chapter 45. Note that for the small driv<strong>in</strong>g amplitude =05<br />

the Lyapunov plot converges to ordered motion with an exponent = −0251 whereas for =12 the<br />

plot diverges characteristic of chaotic motion with an exponent =01538 The Lyapunov exponent usually<br />

fluctuates widely at the local oscillator frequency, and thus the time average of the Lyapunov exponent must<br />

be taken over many periods of the oscillation to identify the general trend with time. Some systems near an<br />

order-to-chaos transition can exhibit positive Lyapunov exponents for short times, characteristic of chaos,<br />

and then converge to negative at longer time imply<strong>in</strong>g ordered motion. The Lyapunov exponents are<br />

used extensively to monitor the stability of the solutions for non-l<strong>in</strong>ear systems. For example the Lyapunov<br />

exponent is used to identify whether fluid flow is lam<strong>in</strong>ar or turbulent as discussed <strong>in</strong> chapter 168.<br />

A dynamical system <strong>in</strong> -dimensional phase space will have a set of Lyapunov exponents { 1 2 }<br />

associated with a set of attractors, the importance of which depend on the <strong>in</strong>itial conditions. Typically one<br />

Lyapunov exponent dom<strong>in</strong>ates at one specific location <strong>in</strong> phase space, and thus it is usual to use the maximal<br />

Lyapunov exponent to identify chaos.The Lyapunov exponent is a very sensitive measure of the onset of chaos<br />

and provides an important test of the chaotic nature for the complicated motion exhibited by non-l<strong>in</strong>ear<br />

systems.<br />

4.6.2 Bifurcation diagram<br />

The bifurcation diagram simplifies the presentation of the dynamical motion by sampl<strong>in</strong>g the status of<br />

thesystemonceperperiod,synchronizedtothedriv<strong>in</strong>gfrequency, for many sets of <strong>in</strong>itial conditions. The<br />

results are presented graphically as a function of one parameter of the system <strong>in</strong> the bifurcation diagram. For<br />

example, the wildly different behavior <strong>in</strong> the driven damped plane pendulum is represented on a bifurcation<br />

diagram <strong>in</strong> figure 411, which shows the observed angular velocity of the pendulum sampled once per drive<br />

cycle plotted versus drive strength. The bifurcation diagram is obta<strong>in</strong>ed by sampl<strong>in</strong>g either the angle ,<br />

or angular velocity once per drive cycle, that is, it represents the observables of the pendulum us<strong>in</strong>g a<br />

stroboscopic technique that samples the motion synchronous with the drive frequency. Bifurcation plots also<br />

can be created as a function of either the time ˜, thedamp<strong>in</strong>gfactor , the normalized frequency ˜ = 0<br />

,<br />

or the driv<strong>in</strong>g amplitude


100 CHAPTER 4. NONLINEAR SYSTEMS AND CHAOS<br />

In the doma<strong>in</strong> with drive strength <br />

10663 there is one unique angle each drive<br />

cycle as illustrated by the bifurcation diagram.<br />

For slightly higher drive strength<br />

period-two bifurcation behavior results <strong>in</strong><br />

two different angles per drive cycle. The<br />

Lyapunov exponent is negative for this region<br />

correspond<strong>in</strong>g to ordered motion. The<br />

cascade of period doubl<strong>in</strong>g with <strong>in</strong>crease <strong>in</strong><br />

drive strength is readily apparent until chaos<br />

sets <strong>in</strong> at the critical drive strength when<br />

there is a random distribution of sampled angular<br />

velocities and the Lyapunov exponent<br />

becomes positive. Note that at =10845<br />

there is a brief <strong>in</strong>terval of period-6 motion<br />

followed by another region of chaos. Around<br />

=11 there is a region that is primarily<br />

chaotic which is reflected by chaotic values of<br />

the angular velocity on the bifurcation plot<br />

and large positive values of the Lyapunov exponent.<br />

The region around =112 exhibits<br />

period three motion and negative Lyapunov<br />

exponent correspond<strong>in</strong>g to ordered motion.<br />

The 115 125 region is ma<strong>in</strong>ly chaotic<br />

and has a large positive Lyapunov exponent.<br />

The region with 13 14 is strik<strong>in</strong>g<br />

<strong>in</strong> that this corresponds to roll<strong>in</strong>g motion<br />

with reemergence of period one and negative<br />

Figure 4.11: Bifurcation diagram samples the angular velocity<br />

once per period for the driven, l<strong>in</strong>early-damped, plane pendulum<br />

plotted as a function of the drive strength . Regions<br />

of period doubl<strong>in</strong>g, and chaos, as well as islands of stability<br />

all are manifest as the drive strength is changed. Note that<br />

the limited number of samples causes broaden<strong>in</strong>g of the l<strong>in</strong>es<br />

adjacent to bifurcations.<br />

Lyapunov exponent. This period-1 motion<br />

is due to a cont<strong>in</strong>uous roll<strong>in</strong>g motion of the<br />

plane pendulum as shown <strong>in</strong> figure 47 where it is seen that the average <strong>in</strong>creases 2 per cycle, whereas the<br />

angular velocity exhibits a periodic motion. That is, on average the pendulum is rotat<strong>in</strong>g 2 per cycle.<br />

Above =14 the system start to exhibit period doubl<strong>in</strong>g followed by chaos rem<strong>in</strong>iscent of the behavior<br />

seen at lower values.<br />

These results show that the bifurcation diagram nicely illustrates the order to chaos transitions for the<br />

harmonically-driven, l<strong>in</strong>early-damped, pendulum. Several transitions between order and chaos are seen to<br />

occur. The apparent ordered and chaotic regimes are confirmed by the correspond<strong>in</strong>g Lyapunov exponents<br />

which alternate between negative and positive values for the ordered and chaotic regions respectively.<br />

4.6.3 Po<strong>in</strong>caré Section<br />

State-space plots are very useful for characteriz<strong>in</strong>g periodic motion, but they become too dense for useful<br />

<strong>in</strong>terpretation when the system approaches chaos as illustrated <strong>in</strong> figure 411 Po<strong>in</strong>caré sections solve this<br />

difficulty by tak<strong>in</strong>g a stroboscopic sample once per cycle of the state-space diagram. That is, the po<strong>in</strong>t on<br />

the state space orbit is sampled once per drive frequency. For period-1 motion this corresponds to a s<strong>in</strong>gle<br />

po<strong>in</strong>t ( ). For period-2 motion this corresponds to two po<strong>in</strong>ts etc. For chaotic systems the sequence of<br />

state-space sample po<strong>in</strong>ts follow complicated trajectories. Figure 412 shows the Po<strong>in</strong>caré sections for the<br />

correspond<strong>in</strong>g state space diagram shown <strong>in</strong> figure 49 for cycles 10 to 6000. Note the complicated curves do<br />

not cross or repeat. Enlargements of any part of this plot will show <strong>in</strong>creas<strong>in</strong>gly dense parallel trajectories,<br />

called fractals, that <strong>in</strong>dicates the complexity of the chaotic cyclic motion. That is, zoom<strong>in</strong>g <strong>in</strong> on a small<br />

section of this Po<strong>in</strong>caré plot shows many closely parallel trajectories. The fractal attractors are surpris<strong>in</strong>gly<br />

robust to large differences <strong>in</strong> <strong>in</strong>itial conditions. Po<strong>in</strong>caré sections are a sensitive probe of periodic motion<br />

for systems where periodic motion is not readily apparent.<br />

In summary, the behavior of the well-known, harmonically-driven, l<strong>in</strong>early-damped, plane pendulum<br />

becomes remarkably complicated at large driv<strong>in</strong>g amplitudes where non-l<strong>in</strong>ear effects dom<strong>in</strong>ate. That is,


4.7. WAVE PROPAGATION FOR NON-LINEAR SYSTEMS 101<br />

Figure 4.12: Three Po<strong>in</strong>caré section plots for the harmonically-driven, l<strong>in</strong>early-damped, pendulum for various<br />

<strong>in</strong>itial conditions with =12 ˜ = 2 3 and ∆ =<br />

<br />

100<br />

. These calculations used the Runge-Kutta method<br />

and were performed for 6000 by E. Shah (Private communication).<br />

whentherestor<strong>in</strong>gforceisnon-l<strong>in</strong>ear. Thesystemexhibits bifurcation where it can evolve to multiple<br />

attractors that depend sensitively on the <strong>in</strong>itial conditions. The system exhibits both oscillatory, and roll<strong>in</strong>g,<br />

solutions depend<strong>in</strong>g on the amplitude of the motion. The system exhibits doma<strong>in</strong>s of simple ordered motion<br />

separated by doma<strong>in</strong>s of very complicated ordered motion as well as chaotic regions. The transitions between<br />

these dramatically different modes of motion are extremely sensitive to the amplitude and phase of the<br />

driver. Eventually the motion becomes completely chaotic. The Lyapunov exponent, bifurcation diagram,<br />

and Po<strong>in</strong>caré section plots, are sensitive measures of the order of the motion. These three sensitive measures<br />

of order and chaos are used extensively <strong>in</strong> many fields <strong>in</strong> classical mechanics. Considerable comput<strong>in</strong>g<br />

capabilities are required to elucidate the complicated motion <strong>in</strong>volved <strong>in</strong> non-l<strong>in</strong>ear systems. Examples<br />

<strong>in</strong>clude lam<strong>in</strong>ar and turbulent flow <strong>in</strong> fluid dynamics and weather forecast<strong>in</strong>g of hurricanes, where the<br />

motion can span a wide dynamic range <strong>in</strong> dimensions from 10 −5 to 10 4 .<br />

4.7 Wave propagation for non-l<strong>in</strong>ear systems<br />

4.7.1 Phase, group, and signal velocities<br />

Chapter 3 discussed the wave equation and solutions for l<strong>in</strong>ear systems. It was shown that, for l<strong>in</strong>ear systems,<br />

the wave motion obeys superposition and exhibits dispersion, that is, a frequency-dependent phase velocity,<br />

and, <strong>in</strong> some cases, attenuation. Nonl<strong>in</strong>ear systems <strong>in</strong>troduce <strong>in</strong>trigu<strong>in</strong>g new wave phenomena. For example<br />

for nonl<strong>in</strong>ear systems, second, and higher terms must be <strong>in</strong>cluded <strong>in</strong> the Taylor expansion given <strong>in</strong> equation<br />

42 These second and higher order terms result <strong>in</strong> the group velocity be<strong>in</strong>g a function of that is, group<br />

velocity dispersion occurs which leads to the shape of the envelope of the wave packet be<strong>in</strong>g time dependent.<br />

As a consequence the group velocity <strong>in</strong> the wave packet is not well def<strong>in</strong>ed, and does not equal the signal<br />

velocity of the wave packet or the phase velocity of the wavelets. Nonl<strong>in</strong>ear optical systems have been studied<br />

experimentally where , which is called slow light, while other systems have which is<br />

called superlum<strong>in</strong>al light. The ability to control the velocity of light <strong>in</strong> such optical systems is of considerable<br />

current <strong>in</strong>terest s<strong>in</strong>ce it has signal transmission applications.<br />

The dispersion relation for a nonl<strong>in</strong>ear system can be expressed as a Taylor expansion of the form<br />

µ <br />

( − 0 )+ 1 <br />

= 0<br />

2<br />

µ 2 <br />

2 <br />

= 0 +<br />

( − 0 ) 2 + (4.41)<br />

= 0<br />

where is used as the <strong>in</strong>dependent variable s<strong>in</strong>ce it is <strong>in</strong>variant to phase transitions of the system. Note<br />

that the factor for the first derivative term is the reciprocal of the group velocity<br />

µ <br />

<br />

= 0<br />

≡ 1<br />

<br />

(4.42)


102 CHAPTER 4. NONLINEAR SYSTEMS AND CHAOS<br />

while the factor for the second derivative term is<br />

µ 2 <br />

∙<br />

<br />

2<br />

µ<br />

− 1<br />

2 <br />

= <br />

¸<br />

<br />

1<br />

<br />

=<br />

(4.43)<br />

= 0<br />

()<br />

= 0<br />

<br />

= 0<br />

which gives the velocity dispersion for the system.<br />

S<strong>in</strong>ce<br />

=<br />

<br />

(4.44)<br />

<br />

then<br />

<br />

≡ 1 = 1 + 1<br />

<br />

(4.45)<br />

<br />

The <strong>in</strong>verse velocities for electromagnetic waves are best represented <strong>in</strong> terms of the correspond<strong>in</strong>g refractive<br />

<strong>in</strong>dices where<br />

≡<br />

<br />

<br />

(4.46)<br />

and the group refractive <strong>in</strong>dex<br />

≡<br />

<br />

(4.47)<br />

<br />

Then equation 445 canbewritten<strong>in</strong>themoreconvenientform<br />

= + <br />

(4.48)<br />

<br />

Wave propagation for an optical system that<br />

is subject to a s<strong>in</strong>gle resonance gives one example<br />

of nonl<strong>in</strong>ear frequency response that has<br />

applications to optics.<br />

Figure 413 shows that the real and imag<strong>in</strong>ary<br />

parts of the phase refractive <strong>in</strong>dex exhibit<br />

the characteristic resonance frequency dependence<br />

of the s<strong>in</strong>usoidally-driven, l<strong>in</strong>ear oscillator<br />

that was discussed <strong>in</strong> chapter 36 and as<br />

illustrated <strong>in</strong> figure 310. Figure413 also shows<br />

the group refractive <strong>in</strong>dex computed us<strong>in</strong>g<br />

equation 448.<br />

Note that at resonance, is reduced below<br />

the non-resonant value which corresponds<br />

to superlum<strong>in</strong>al (fast) light, whereas <strong>in</strong> the<br />

w<strong>in</strong>gs of the resonance is larger than the<br />

non-resonant value correspond<strong>in</strong>g to slow light.<br />

Thus the nonl<strong>in</strong>ear dependence of the refractive<br />

<strong>in</strong>dex on angular frequency leads to fast<br />

or slow group velocities for isolated wave packets.<br />

Velocities of light as slow as 17 sec have<br />

been observed. Experimentally the energy absorption<br />

that occurs on resonance makes it difficult<br />

to observe the superlum<strong>in</strong>al electromagnetic<br />

wave at resonance.<br />

Note that Sommerfeld and Brillou<strong>in</strong> showed<br />

that even though the group velocity may exceed<br />

, the signal velocity, which marks the arrival of<br />

the lead<strong>in</strong>g edge of the optical pulse, does not<br />

exceed , the velocity of light <strong>in</strong> vacuum, as was<br />

postulated by E<strong>in</strong>ste<strong>in</strong>.[Bri14]<br />

Figure 4.13: The real and imag<strong>in</strong>ary parts of the phase<br />

refractive <strong>in</strong>dex n plus the real part of the group refractive<br />

<strong>in</strong>dex associated with an isolated atomic resonance.


4.7. WAVE PROPAGATION FOR NON-LINEAR SYSTEMS 103<br />

4.7.2 Soliton wave propagation<br />

The soliton is a fasc<strong>in</strong>at<strong>in</strong>g and very special<br />

wave propagation phenomenon that occurs for<br />

certa<strong>in</strong> non-l<strong>in</strong>ear systems. The soliton is a selfre<strong>in</strong>forc<strong>in</strong>g<br />

solitary localized wave packet that<br />

ma<strong>in</strong>ta<strong>in</strong>s its shape while travell<strong>in</strong>g long distances<br />

at a constant speed. Solitons are caused by a<br />

cancellation of phase modulation result<strong>in</strong>g from<br />

non-l<strong>in</strong>ear velocity dependence, and the group velocity<br />

dispersive effects <strong>in</strong> a medium. Solitons<br />

arise as solutions of a widespread class of weaklynonl<strong>in</strong>ear<br />

dispersive partial differential equations<br />

describ<strong>in</strong>g many physical systems. Figure 414<br />

shows a soliton compris<strong>in</strong>g a solitary water wave<br />

approach<strong>in</strong>g the coast of Hawaii. While the soliton<br />

<strong>in</strong> Fig. 414 may appear like a normal wave,<br />

it is unique <strong>in</strong> that there are no other waves accompany<strong>in</strong>g<br />

it. This wave was probably created<br />

far away from the shore when a normal wave was<br />

modulated by a geometrical change <strong>in</strong> the ocean<br />

depth, such as the ris<strong>in</strong>g sea floor, which forced<br />

it <strong>in</strong>to the appropriate shape for a soliton. The<br />

wave then was able to travel to the coast <strong>in</strong>tact,<br />

Figure 4.14: A solitary wave approaches the coast of Hawaii.<br />

(Image: Robert Odom/University of Wash<strong>in</strong>gton)<br />

despite the apparently placid nature of the ocean near the beach. Solitons are notable <strong>in</strong> that they <strong>in</strong>teract<br />

with each other <strong>in</strong> ways very different from normal waves. Normal waves are known for their complicated<br />

<strong>in</strong>terference patterns that depend on the frequency and wavelength of the waves. Solitons, can pass right<br />

through each other without be<strong>in</strong>g a affected at all. This makes solitons very appeal<strong>in</strong>g to scientists because<br />

soliton waves are more sturdy than normal waves, and can therefore be used to transmit <strong>in</strong>formation <strong>in</strong> ways<br />

that are dist<strong>in</strong>ctly different than for normal wave motion. For example, optical solitons are used <strong>in</strong> optical<br />

fibers made of a dispersive, nonl<strong>in</strong>ear optical medium, to transmit optical pulses with an <strong>in</strong>variant shape.<br />

Solitons were first observed <strong>in</strong> 1834 by John Scott Russell (1808 − 1882). Russell was an eng<strong>in</strong>eer conduct<strong>in</strong>g<br />

experiments to <strong>in</strong>crease the efficiency of canal boats. His experimental and theoretical <strong>in</strong>vestigations<br />

allowedhimtorecreatethephenomenon<strong>in</strong>wavetanks. Through his extensive studies, Scott Russell noticed<br />

that soliton propagation exhibited the follow<strong>in</strong>g properties:<br />

• The waves are stable and hold their shape for long periods of time.<br />

• The waves can travel over long distances at uniform speed.<br />

• The speed of propagation of the wave depends on the size of the wave, with larger waves travel<strong>in</strong>g<br />

faster than smaller waves.<br />

• The waves ma<strong>in</strong>ta<strong>in</strong>ed their shape when they collided - seem<strong>in</strong>gly pass<strong>in</strong>g right through each other.<br />

Scott Russell’s work was met with scepticism by the scientific community. The problem with the Wave<br />

of Translation was that it was an effect that depended on nonl<strong>in</strong>ear effects, whereas previously exist<strong>in</strong>g<br />

theories of hydrodynamics (such as those of Newton and Bernoulli) only dealt with l<strong>in</strong>ear systems. George<br />

Biddell Airy, and George Gabriel Stokes, published papers attack<strong>in</strong>g Scott Russell’s observations because<br />

the observations could not be expla<strong>in</strong>ed by their theories of wave propagation <strong>in</strong> water. Regardless, Scott<br />

Russell was conv<strong>in</strong>ced of the prime importance of the Wave of Translation, and history proved that he was<br />

correct. Scott Russell went on to develop the “wave l<strong>in</strong>e” system of hull construction that revolutionized<br />

n<strong>in</strong>eteenth century naval architecture, along with a number of other great accomplishments lead<strong>in</strong>g him to<br />

fame and prom<strong>in</strong>ence. Despite all of the success <strong>in</strong> his career, he cont<strong>in</strong>ued throughout his life to pursue his<br />

studies of the Wave of Translation.<br />

In 1895 Korteweg and de Vries developed a wave equation for surface waves for shallow water.<br />

<br />

+ 3 <br />

3 +6 =0<br />

<br />

(4.49)<br />

A solution of this equation has the characteristics of a solitary wave with fixed shape. It is given by<br />

substitut<strong>in</strong>g the form ( ) =( − ) <strong>in</strong>to the Korteweg-de Vries equation which gives


104 CHAPTER 4. NONLINEAR SYSTEMS AND CHAOS<br />

− <br />

+ 3 <br />

+6 =0 (4.50)<br />

3 <br />

Integrat<strong>in</strong>g with respect to gives<br />

3 2 + 2 <br />

− = (4.51)<br />

3 where is a constant of <strong>in</strong>tegration. This non-l<strong>in</strong>ear equation has a solution<br />

( ) = 1 ∙√ <br />

2 sec 2 ( − − )¸<br />

(4.52)<br />

2<br />

where is a constant. Equation 452 is the equation of a solitary wave mov<strong>in</strong>g <strong>in</strong> the + direction at a<br />

velocity .<br />

Soliton behavior is observed <strong>in</strong> phenomena such as tsunamis, tidal bores that occur for some rivers,<br />

signals <strong>in</strong> optical fibres, plasmas, atmospheric waves, vortex filaments, superconductivity, and gravitational<br />

fields hav<strong>in</strong>g cyl<strong>in</strong>drical symmetry. Much work has been done on solitons for fibre optics applications. The<br />

soliton’s <strong>in</strong>herent stability make long-distance transmission possible without the use of repeaters, and could<br />

potentially double the transmission capacity.<br />

Before the discovery of solitons, mathematicians were under the impression that nonl<strong>in</strong>ear partial differential<br />

equations could not be solved exactly. However, solitons led to the recognition that there are non-l<strong>in</strong>ear<br />

systems that can be solved analytically. This discovery haspromptedmuch<strong>in</strong>vestigation<strong>in</strong>totheseso-called<br />

“<strong>in</strong>tegrable systems.” Such systems are rare, as most non-l<strong>in</strong>ear differential equations admit chaotic behavior<br />

with no explicit solutions. Integrable systems nevertheless lead to very <strong>in</strong>terest<strong>in</strong>g mathematics rang<strong>in</strong>g from<br />

differential geometry and complex analysis to quantum field theory and fluid dynamics.<br />

Many of the fundamental equations <strong>in</strong> physics (Maxwell’s, Schröd<strong>in</strong>ger’s) are l<strong>in</strong>ear equations. However,<br />

physicists have begun to recognize many areas of physics <strong>in</strong> which nonl<strong>in</strong>earity can result <strong>in</strong> qualitatively<br />

new phenomenon which cannot be constructed via perturbation theory start<strong>in</strong>g from l<strong>in</strong>earized equations.<br />

These <strong>in</strong>clude phenomena <strong>in</strong> magnetohydrodynamics, meteorology, oceanography, condensed matter physics,<br />

nonl<strong>in</strong>ear optics, and elementary particle physics. For example, the European space mission Cluster detected<br />

a soliton-like electrical disturbances that travelled through the ionized gas surround<strong>in</strong>g the Earth start<strong>in</strong>g<br />

about 50,000 kilometers from Earth and travell<strong>in</strong>g towards the planet at about 8 km/s. It is thought that<br />

this soliton was generated by turbulence <strong>in</strong> the magnetosphere.<br />

Efforts to understand the nonl<strong>in</strong>earity of solitons has led to much research <strong>in</strong> many areas of physics. In<br />

the context of solitons, their particle-like behavior (<strong>in</strong> that they are localized and preserved under collisions)<br />

leads to a number of experimental and theoretical applications. The technique known as bosonization allows<br />

view<strong>in</strong>g particles, such as electrons and positrons, as solitons <strong>in</strong> appropriate field equations. There are<br />

numerous macroscopic phenomena, such as <strong>in</strong>ternal waves on the ocean, spontaneous transparency, and the<br />

behavior of light <strong>in</strong> fiber optic cable, that are now understood <strong>in</strong> terms of solitons. These phenomena are<br />

be<strong>in</strong>g applied to modern technology.<br />

4.8 Summary<br />

The study of the dynamics of non-l<strong>in</strong>ear systems rema<strong>in</strong>s a vibrant and rapidly evolv<strong>in</strong>g field<strong>in</strong>classical<br />

mechanics as well as many other branches of science. This chapter has discussed examples of non-l<strong>in</strong>ear<br />

systems <strong>in</strong> classical mechanics. It was shown that the superposition pr<strong>in</strong>ciple is broken even for weak<br />

nonl<strong>in</strong>earity. It was shown that <strong>in</strong>creased nonl<strong>in</strong>earity leads to bifurcation, po<strong>in</strong>t attractors, limit-cycle<br />

attractors, and sensitivity to <strong>in</strong>itial conditions.<br />

Limit-cycle attractors: The Po<strong>in</strong>caré-Bendixson theorem for limit cycle attractors states that the<br />

paths, both <strong>in</strong> state-space and phase-space, can have three possible paths:<br />

(1) closed paths, like the elliptical paths for the undamped harmonic oscillator,<br />

(2) term<strong>in</strong>ate at an equilibrium po<strong>in</strong>t as →∞, like the po<strong>in</strong>t attractor for a damped harmonic oscillator,<br />

(3) tend to a limit cycle as →∞.<br />

The limit cycle is unusual <strong>in</strong> that the periodic motion tends asymptotically to the limit-cycle attractor<br />

<strong>in</strong>dependent of whether the <strong>in</strong>itial values are <strong>in</strong>side or outside the limit cycle. The balance of dissipative forces<br />

and driv<strong>in</strong>g forces often leads to limit-cycle attractors, especially <strong>in</strong> biological applications. Identification of


4.8. SUMMARY 105<br />

limit-cycle attractors, as well as the trajectories of the motion towards these limit-cycle attractors, is more<br />

complicated than for po<strong>in</strong>t attractors.<br />

The van der Pol oscillator is a common example of a limit-cycle system that has an equation of motion<br />

of the form<br />

2 <br />

2 + ¡ 2 − 1 ¢ <br />

+ 2 0 =0<br />

(411)<br />

The van der Pol oscillator has a limit-cycle attractor that <strong>in</strong>cludes non-l<strong>in</strong>ear damp<strong>in</strong>g and exhibits<br />

periodic solutions that asymptotically approach one attractor solution <strong>in</strong>dependent of the <strong>in</strong>itial conditions.<br />

There are many examples <strong>in</strong> nature that exhibit similar behavior.<br />

Harmonically-driven, l<strong>in</strong>early-damped, plane pendulum: The non-l<strong>in</strong>earity of the well-known<br />

driven l<strong>in</strong>early-damped plane pendulum was used as an example of the behavior of non-l<strong>in</strong>ear systems <strong>in</strong><br />

nature. It was shown that non-l<strong>in</strong>earity leads to discont<strong>in</strong>uous period bifurcation, extreme sensitivity to<br />

<strong>in</strong>itial conditions, roll<strong>in</strong>g motion and chaos.<br />

Differentiation between ordered and chaotic motion: Lyapunov exponents, bifurcation diagrams,<br />

and Po<strong>in</strong>caré sections were used to identify the transition from order to chaos. Chapter 168 discusses<br />

the non-l<strong>in</strong>ear Navier-Stokes equations of viscous-fluid flow which leads to complicated transitions between<br />

lam<strong>in</strong>ar and turbulent flow. Fluid flow exhibits remarkable complexity that nicely illustrates the dom<strong>in</strong>ant<br />

role that non-l<strong>in</strong>earity can have on the solutions of practical non-l<strong>in</strong>ear systems <strong>in</strong> classical mechanics.<br />

Wave propagation for non-l<strong>in</strong>ear systems: Non-l<strong>in</strong>ear equations can lead to unexpected behavior<br />

for wave packet propagation such as fast or slow light as well as soliton solutions. Moreover, it is notable<br />

that some non-l<strong>in</strong>ear systems can lead to analytic solutions.<br />

The complicated phenomena exhibited by the above non-l<strong>in</strong>ear systems is not restricted to classical<br />

mechanics, rather it is a manifestation of the mathematical behavior of the solutions of the differential<br />

equations <strong>in</strong>volved. That is, this behavior is a general manifestation of the behavior of solutions for secondorder<br />

differential equations. Exploration of this complex motion has only become feasible with the advent<br />

of powerful computer facilities dur<strong>in</strong>g the past three decades. The breadth of phenomena exhibited by<br />

these examples is manifest <strong>in</strong> myriads of other nonl<strong>in</strong>ear systems, rang<strong>in</strong>g from many-body motion, weather<br />

patterns, growth of biological species, epidemics, motion of electrons <strong>in</strong> atoms, etc. Other examples of nonl<strong>in</strong>ear<br />

equations of motion not discussed here, are the three-body problem, which is mentioned <strong>in</strong> chapter<br />

11, and turbulence <strong>in</strong> fluid flow which is discussed <strong>in</strong> chapter 16.<br />

It is stressed that the behavior discussed <strong>in</strong> this chapter is very different from the random walk problem<br />

which is a stochastic process where each step is purely random and not determ<strong>in</strong>istic. This chapter has<br />

assumed that the motion is fully determ<strong>in</strong>istic and rigorously follows the laws of classical mechanics. Even<br />

though the motion is fully determ<strong>in</strong>istic, and follows the laws of classical mechanics, the motion is extremely<br />

sensitive to the <strong>in</strong>itial conditions and the non-l<strong>in</strong>earities can lead to chaos. Computer modell<strong>in</strong>g is the only<br />

viable approach for predict<strong>in</strong>g the behavior of such non-l<strong>in</strong>ear systems. The complexity of solv<strong>in</strong>g non-l<strong>in</strong>ear<br />

equations is the reason that this book will cont<strong>in</strong>ue to consider only l<strong>in</strong>ear systems. Fortunately, <strong>in</strong> nature,<br />

non-l<strong>in</strong>ear systems can be approximately l<strong>in</strong>ear when the small-amplitude assumption is applicable.


106 CHAPTER 4. NONLINEAR SYSTEMS AND CHAOS<br />

Problems<br />

1. Consider the chaotic motion of the driven damped pendulum whose equation of motion is given by<br />

¨ + Γ ˙ + 2 0 s<strong>in</strong> = 2 0 cos <br />

for which the Lyapunov exponent is =1with time measured <strong>in</strong> units of the drive period.<br />

(a) Assume that you need to predict () with accuracy of 10 −2 , and that the <strong>in</strong>itial value (0) is<br />

known to with<strong>in</strong> 10 −6 . What is the maximum time horizon max for which you can predict ()<br />

to with<strong>in</strong> the required accuracy?<br />

(b) Suppose that you manage to improve the accuracy of the <strong>in</strong>itial value to 10 −9 (that is, a thousandfold<br />

improvement). What is the time horizon now for achiev<strong>in</strong>g the accuracy of 10 −2 ?<br />

(c) By what factor has max improved with the 1000 − improvement <strong>in</strong> <strong>in</strong>itial measurement.<br />

(d) What does this imply regard<strong>in</strong>g long-term predictions of chaotic motion?<br />

2. A non-l<strong>in</strong>ear oscillator satisfies the equation ¨ + ˙ 3 + =0 F<strong>in</strong>d the polar equations for the motion <strong>in</strong> the<br />

state-space diagram. Show that any trajectory that starts with<strong>in</strong> the circle 1 encircle the orig<strong>in</strong> <strong>in</strong>f<strong>in</strong>itely<br />

many times <strong>in</strong> the clockwise direction. Show further that these trajectories <strong>in</strong> state space term<strong>in</strong>ate at the<br />

orig<strong>in</strong>.<br />

3. Consider the system of a mass suspended between two identical spr<strong>in</strong>gs as shown.<br />

If each spr<strong>in</strong>g is stretched a distance to attach the mass at the equilibrium position the mass is subject to<br />

two equal and oppositely directed forces of magnitude . Ignore gravity. Show that the potential <strong>in</strong> which<br />

the mass moves is approximately<br />

() =<br />

Construct a state-space diagram for this potential.<br />

4. A non-l<strong>in</strong>ear oscillator satisfies the equation<br />

½ ¾ <br />

2 +<br />

<br />

½ ( − )<br />

4 3<br />

¨ +( 2 + ˙ 2 − 1) ˙ + =0<br />

F<strong>in</strong>d the polar equations for the motion <strong>in</strong> the state-space diagram. Show that any trajectory that starts <strong>in</strong><br />

the doma<strong>in</strong> 1 √ 3 spirals clockwise and tends to the limit cycle =1.[Thesameistrueoftrajectories<br />

that start <strong>in</strong> the doma<strong>in</strong> 0 1. ] What is the period of the limit cycle?<br />

5. A mass moves <strong>in</strong> one direction and is subject to a constant force + 0 when 0 and to a constant force<br />

− 0 when 0. Describe the motion by construct<strong>in</strong>g a state space diagram. Calculate the period of the<br />

motion <strong>in</strong> terms of 0 and the amplitude . Disregard damp<strong>in</strong>g.<br />

¾<br />

4<br />

6. Investigate the motion of an undamped mass subject to a force of the form<br />

() =( −<br />

−( + ) + <br />

|| <br />

||


Chapter 5<br />

Calculus of variations<br />

5.1 Introduction<br />

The prior chapters have focussed on the <strong>in</strong>tuitive Newtonian approach to classical mechanics, which is based<br />

on vector quantities like force, momentum, and acceleration. Newtonian mechanics leads to second-order<br />

differential equations of motion. The calculus of variations underlies a powerful alternative approach to<br />

classical mechanics that is based on identify<strong>in</strong>g the path that m<strong>in</strong>imizes an <strong>in</strong>tegral quantity. This <strong>in</strong>tegral<br />

variational approach was first championed by Gottfried Wilhelm Leibniz, contemporaneously with Newton’s<br />

development of the differential approach to classical mechanics.<br />

Dur<strong>in</strong>g the 18 century, Bernoulli, who was a student of Leibniz, developed the field of variational<br />

calculus which underlies the <strong>in</strong>tegral variational approach to mechanics. He solved the brachistochrone<br />

problem which <strong>in</strong>volves f<strong>in</strong>d<strong>in</strong>g the path for which the transit time between two po<strong>in</strong>ts is the shortest. The<br />

<strong>in</strong>tegral variational approach also underlies Fermat’s pr<strong>in</strong>ciple <strong>in</strong> optics, which can be used to derive that<br />

the angle of reflection equals the angle of <strong>in</strong>cidence, as well as derive Snell’s law. Other applications of the<br />

calculus of variations <strong>in</strong>clude solv<strong>in</strong>g the catenary problem, f<strong>in</strong>d<strong>in</strong>g the maximum and m<strong>in</strong>imum distances<br />

between two po<strong>in</strong>ts on a surface, polygon shapes hav<strong>in</strong>g the maximum ratio of enclosed area to perimeter,<br />

or maximiz<strong>in</strong>g profit <strong>in</strong> economics. Bernoulli, developed the pr<strong>in</strong>ciple of virtual work used to describe<br />

equilibrium <strong>in</strong> static systems, and d’Alembert extended the pr<strong>in</strong>ciple of virtual work to dynamical systems.<br />

Euler, the preem<strong>in</strong>ent Swiss mathematician of the 18 century and a student of Bernoulli, developed the<br />

calculus of variations with full mathematical rigor. The culm<strong>in</strong>ation of the development of the Lagrangian<br />

variational approach to classical mechanics is done by Lagrange (1736-1813), who was a student of Euler,.<br />

The Euler-Lagrangian approach to classical mechanics stems from a deep philosophical belief that the<br />

laws of nature are based on the pr<strong>in</strong>ciple of economy.That is, the physical universe follows paths through<br />

space and time that are based on extrema pr<strong>in</strong>ciples. The standard Lagrangian is def<strong>in</strong>ed as the difference<br />

between the k<strong>in</strong>etic and potential energy, that is<br />

= − (5.1)<br />

Chapters 6 through 9 will show that the laws of classical mechanics can be expressed <strong>in</strong> terms of Hamilton’s<br />

variational pr<strong>in</strong>ciple which states that the motion of the system between the <strong>in</strong>itial time 1 and f<strong>in</strong>al time<br />

2 follows a path that m<strong>in</strong>imizes the scalar action <strong>in</strong>tegral def<strong>in</strong>ed as the time <strong>in</strong>tegral of the Lagrangian.<br />

Z 2<br />

= (5.2)<br />

1<br />

The calculus of variations provides the mathematics required to determ<strong>in</strong>e the path that m<strong>in</strong>imizes the<br />

action <strong>in</strong>tegral. This variational approach is both elegant and beautiful, and has withstood the rigors of<br />

experimental confirmation. In fact, not only is it an exceed<strong>in</strong>gly powerful alternative approach to the <strong>in</strong>tuitive<br />

Newtonian approach <strong>in</strong> classical mechanics, but Hamilton’s variational pr<strong>in</strong>ciple now is recognized to be more<br />

fundamental than Newton’s Laws of Motion. The Lagrangian and Hamiltonian variational approaches to<br />

mechanics are the only approaches that can handle the Theory of Relativity, statistical mechanics, and the<br />

dichotomy of philosophical approaches to quantum physics.<br />

107


108 CHAPTER 5. CALCULUS OF VARIATIONS<br />

5.2 Euler’s differential equation<br />

The calculus of variations, presented here, underlies the powerful variational approaches that were developed<br />

for classical mechanics. <strong>Variational</strong> calculus, developed for classical mechanics, now has become an essential<br />

approach to many other discipl<strong>in</strong>es <strong>in</strong> science, eng<strong>in</strong>eer<strong>in</strong>g, economics, and medic<strong>in</strong>e.<br />

For the special case of one dimension, the calculus of variations reduces to vary<strong>in</strong>g the function () such<br />

that the scalar functional is an extremum, that is, it is a maximum or m<strong>in</strong>imum, where.<br />

=<br />

Z 2<br />

1<br />

[() 0 (); ] (5.3)<br />

Here is the <strong>in</strong>dependent variable, () the dependent variable, plus its first derivative 0 ≡ <br />

The quantity<br />

[() 0 (); ] has some given dependence on 0 and The calculus of variations <strong>in</strong>volves vary<strong>in</strong>g the<br />

function () until a stationary value of is found, which is presumed to be an extremum. This means that<br />

if a function = () gives a m<strong>in</strong>imum value for the scalar functional , then any neighbor<strong>in</strong>g function, no<br />

matter how close to () must <strong>in</strong>crease . For all paths, the <strong>in</strong>tegral is taken between two fixed po<strong>in</strong>ts,<br />

1 1 and 2 2 Possible paths between the <strong>in</strong>itial and f<strong>in</strong>al po<strong>in</strong>ts are illustrated <strong>in</strong> figure 51. Relativeto<br />

any neighbor<strong>in</strong>g path, the functional must have a stationary value which is presumed to be the correct<br />

extremum path.<br />

Def<strong>in</strong>e a neighbor<strong>in</strong>g function us<strong>in</strong>g a parametric representation ( ) such that for =0, = (0)=<br />

() is the function that yields the extremum for . Assume that an <strong>in</strong>f<strong>in</strong>itesimally small fraction of the<br />

neighbor<strong>in</strong>g function () is added to the extremum path (). Thatis,assume<br />

( ) = (0)+() (5.4)<br />

0 ( ) ≡<br />

( )<br />

= (0) + <br />

<br />

[( ) ( ); ] (5.5)<br />

whereitisassumedthattheextremumfunction(0) and the auxiliary function () are well behaved<br />

functions of with cont<strong>in</strong>uous first derivatives, and where () vanishes at 1 and 2 because, for all possible<br />

paths, the function ( ) must be identical with () at the end po<strong>in</strong>ts of the path, i.e. ( 1 )=( 2 )=0.<br />

The situation is depicted <strong>in</strong> figure 51. It is possible to express any such parametric family of curves as<br />

afunctionof<br />

Z 2<br />

() =<br />

1<br />

The condition that the <strong>in</strong>tegral has a stationary (extremum) value is that be <strong>in</strong>dependent of to first<br />

order along the path. That is, the extremum value occurs for =0where<br />

µ <br />

<br />

=0 (5.6)<br />

1<br />

=0<br />

for all functions () This is illustrated on the right side of figure 51<br />

Apply<strong>in</strong>g condition (56) to equation (55) and s<strong>in</strong>ce is <strong>in</strong>dependent of then<br />

Z<br />

<br />

2<br />

µ <br />

= <br />

+ 0 <br />

0 =0 (5.7)<br />

<br />

S<strong>in</strong>ce the limits of <strong>in</strong>tegration are fixed, the differential operation affects only the <strong>in</strong>tegrand. From equations<br />

(54),<br />

<br />

= () (5.8)<br />

<br />

and<br />

Consider the second term <strong>in</strong> the <strong>in</strong>tegrand<br />

Z 2<br />

1<br />

0<br />

= <br />

<br />

0 Z 2<br />

0 =<br />

1<br />

(5.9)<br />

<br />

0 (5.10)


5.2. EULER’S DIFFERENTIAL EQUATION 109<br />

y(x)<br />

Varied path<br />

Extremum path, y(x)<br />

x 1 x 2<br />

x<br />

x<br />

O<br />

Figure 5.1: The left shows the extremum () and neighbor<strong>in</strong>g paths ( ) =()+() between ( 1 1 )<br />

and ( 2 2 ) that m<strong>in</strong>imizes the function = R 2<br />

1<br />

[() 0 (); ] . The right shows the dependence of <br />

as a function of the admixture coefficient for a maximum (upper) or a m<strong>in</strong>imum (lower) at =0.<br />

Integrate by parts<br />

Z<br />

Z<br />

= −<br />

(5.11)<br />

gives<br />

Z 2<br />

∙ Z<br />

<br />

()¸2 2<br />

1<br />

0 = 0 − () µ <br />

1 1<br />

0 (5.12)<br />

Note that the first term on the right-hand side is zero s<strong>in</strong>ce by def<strong>in</strong>ition <br />

= () =0at 1 and 2 Thus<br />

<br />

= Z 2<br />

1<br />

Thus equation 57 reduces to<br />

The function <br />

<br />

µ <br />

+ 0 Z 2<br />

µ <br />

<br />

0 = () − ()<br />

<br />

1<br />

<br />

<br />

= Z 2<br />

1<br />

µ <br />

− <br />

<br />

will be an extremum if it is stationary at =0.Thatis,<br />

<br />

= Z 2<br />

1<br />

µ <br />

− <br />

<br />

µ <br />

0 <br />

<br />

<br />

<br />

0 () (5.13)<br />

<br />

<br />

0 () =0 (5.14)<br />

This <strong>in</strong>tegral now appears to be <strong>in</strong>dependent of However, the functions and 0 occurr<strong>in</strong>g <strong>in</strong> the derivatives<br />

are functions of S<strong>in</strong>ce ¡ ¢<br />

<br />

<br />

must vanish for a stationary value, and because () is an arbitrary function<br />

=0<br />

subject to the conditions stated, then the above <strong>in</strong>tegrand must be zero. This derivation that the <strong>in</strong>tegrand<br />

must be zero leads to Euler’s differential equation<br />

<br />

− <br />

<br />

<br />

=0 (5.15)<br />

0 where and 0 are the orig<strong>in</strong>al functions, <strong>in</strong>dependent of The basis of the calculus of variations is that the<br />

function () that satisfies Euler’s equation is an stationary function. Note that the stationary value could<br />

be either a maximum or a m<strong>in</strong>imum value. When Euler’s equation is applied to mechanical systems us<strong>in</strong>g<br />

the Lagrangian as the functional, then Euler’s differential equation is called the Euler-Lagrange equation.


110 CHAPTER 5. CALCULUS OF VARIATIONS<br />

5.3 Applications of Euler’s equation<br />

5.1 Example: Shortest distance between two po<strong>in</strong>ts<br />

Consider the path lies <strong>in</strong> the x − y plane. The <strong>in</strong>f<strong>in</strong>itessimal length of arc is<br />

⎡<br />

= p s ⎤<br />

µ 2<br />

2 + 2 = ⎣ <br />

1+ ⎦ <br />

<br />

Then the length of the arc is<br />

The function is<br />

=<br />

q<br />

1+( 0 ) 2<br />

=<br />

Z 2<br />

1<br />

=<br />

Z 2<br />

1<br />

⎡s<br />

⎣ 1+<br />

y<br />

⎤<br />

µ 2 <br />

⎦ <br />

<br />

Therefore<br />

and<br />

<br />

=0<br />

<br />

0 = 0<br />

q<br />

1+( 0 ) 2<br />

x 1<br />

y 1<br />

x 2<br />

y 2<br />

Insert<strong>in</strong>g these <strong>in</strong>to Euler’s equation 515 gives<br />

⎛ ⎞<br />

0+ ⎝<br />

<br />

q<br />

0<br />

⎠ =0<br />

<br />

1+( 0 ) 2<br />

Shortest distance between two po<strong>in</strong>ts <strong>in</strong> a plane.<br />

x<br />

that is<br />

0<br />

q1+( 0 ) 2 = constant = <br />

This is valid if<br />

0 =<br />

<br />

√<br />

1 − <br />

2 = <br />

Therefore<br />

= + <br />

which is the equation of a straight l<strong>in</strong>e <strong>in</strong> the plane. Thus the shortest path between two po<strong>in</strong>ts <strong>in</strong> a plane is<br />

a straight l<strong>in</strong>e between these po<strong>in</strong>ts, as is <strong>in</strong>tuitively obvious. This stationary value obviously is a m<strong>in</strong>imum.<br />

This trivial example of the use of Euler’s equation to determ<strong>in</strong>e an extremum value has given the obvious<br />

answer. It has been presented here because it provides a proof that a straight l<strong>in</strong>e is the shortest distance <strong>in</strong><br />

a plane and illustrates the power of the calculus of variations to determ<strong>in</strong>e extremum paths.<br />

5.2 Example: Brachistochrone problem<br />

The Brachistochrone problem <strong>in</strong>volves f<strong>in</strong>d<strong>in</strong>g the path hav<strong>in</strong>g the m<strong>in</strong>imum transit time between two<br />

po<strong>in</strong>ts. The Brachistochrone problem stimulated the development of the calculus of variations by John<br />

Bernoulli and Euler. For simplicity, take the case of frictionless motion <strong>in</strong> the − planewithauniform<br />

gravitational field act<strong>in</strong>g <strong>in</strong> the by direction, as shown <strong>in</strong> the adjacent figure. The question is what<br />

constra<strong>in</strong>ed path will result <strong>in</strong> the m<strong>in</strong>imum transit time between two po<strong>in</strong>ts ( 1 1 ) and ( 2 2 )


5.3. APPLICATIONS OF EULER’S EQUATION 111<br />

Consider that the particle of mass starts at the orig<strong>in</strong> 1 =0 1 =0with zero velocity. S<strong>in</strong>ce the<br />

problem conserves energy and assum<strong>in</strong>g that <strong>in</strong>itially = + =0then<br />

That is<br />

The transit time is given by<br />

=<br />

Z 2<br />

1<br />

1<br />

2 2 − =0<br />

= p 2<br />

Z<br />

<br />

2<br />

p<br />

= 2 + 2 Z<br />

s<br />

2<br />

(1 + <br />

√ =<br />

02 )<br />

<br />

1 2 1<br />

2<br />

where 0 ≡ <br />

<br />

. Note that, <strong>in</strong> this example, the <strong>in</strong>dependent variable has been chosen to be and the dependent<br />

variable is ().<br />

The function of the <strong>in</strong>tegral is<br />

s<br />

= √ 1 (1 + 02 )<br />

2 <br />

Factor out the constant √ 2 term, which does not affect the f<strong>in</strong>al equation, and note that<br />

Therefore Euler’s equation gives<br />

⎛<br />

⎞<br />

0+ ⎜<br />

0<br />

⎝<br />

r<br />

<br />

³1+( 2´ ⎟<br />

⎠ =0<br />

0 )<br />

<br />

= 0<br />

<br />

0 =<br />

0<br />

r<br />

<br />

³1+( 2´<br />

0 )<br />

or<br />

(x 1<br />

, y )<br />

a<br />

a<br />

x<br />

1<br />

(x , y )<br />

That is<br />

0<br />

1<br />

r ³ = constant = √<br />

1+( 0 )<br />

2´ 2<br />

02<br />

³<br />

2´ = 1<br />

1+( 0 ) 2<br />

a<br />

2a<br />

P(x , y)<br />

Cycloid<br />

2<br />

2<br />

This may be rewritten as<br />

=<br />

Z 2<br />

1<br />

<br />

p<br />

2 − <br />

2<br />

Change the variable to = (1 − cos ) gives<br />

that = s<strong>in</strong> lead<strong>in</strong>g to the <strong>in</strong>tegral<br />

Z<br />

= (1 − cos ) <br />

y<br />

The Bachistochrone problem <strong>in</strong>volves f<strong>in</strong>d<strong>in</strong>g the path for<br />

the m<strong>in</strong>imum transit time for constra<strong>in</strong>ed frictionless<br />

motion <strong>in</strong> a uniform gravitational field.<br />

or<br />

= ( − s<strong>in</strong> )+constant


112 CHAPTER 5. CALCULUS OF VARIATIONS<br />

The parametric equations for a cycloid pass<strong>in</strong>g through the orig<strong>in</strong> are<br />

= ( − s<strong>in</strong> )<br />

= (1 − cos )<br />

which is the form of the solution found. That is, the shortest time between two po<strong>in</strong>ts is obta<strong>in</strong>ed by constra<strong>in</strong><strong>in</strong>g<br />

the motion of the mass to follow a cycloid shape. Thus the mass first accelerates rapidly by fall<strong>in</strong>g<br />

down steeply and then follows the curve and coasts upward at the end. The elapsed time is obta<strong>in</strong>ed by<br />

<strong>in</strong>sert<strong>in</strong>g q the above parametric relations for and <strong>in</strong> terms of <strong>in</strong>to the transit time <strong>in</strong>tegral giv<strong>in</strong>g<br />

= <br />

<br />

where and are fixed by the end po<strong>in</strong>t coord<strong>in</strong>ates. Thus the time to fall from start<strong>in</strong>g with zero<br />

velocity at the cusp to the m<strong>in</strong>imum of the cycloid is q<br />

<br />

If 2 = 1 =0then 2 =2 which def<strong>in</strong>es the<br />

q<br />

shape of the cycloid and the m<strong>in</strong>imum time is 2q<br />

<br />

= 22<br />

<br />

If the mass starts with a non-zero <strong>in</strong>itial<br />

velocity, then the start<strong>in</strong>g po<strong>in</strong>t is not at the cusp of the cycloid, but down a distance such that the k<strong>in</strong>etic<br />

energy equals the potential energy difference from the cusp.<br />

A modern application of the Brachistochrone problem is determ<strong>in</strong>ation of the optimum shape of the lowfriction<br />

emergency chute that passengers slide down to evacuate a burn<strong>in</strong>g aircraft. Bernoulli solved the<br />

problem of rapid evacuation of an aircraft two centuries before the first flight of a powered aircraft.<br />

5.3 Example: M<strong>in</strong>imal travel cost<br />

Assume that the cost of fly<strong>in</strong>g an aircraft at height is − perunitdistanceofflight-path, where is a<br />

positive constant. Consider that the aircraft flies <strong>in</strong> the ( )-plane from the po<strong>in</strong>t (− 0) to the po<strong>in</strong>t ( 0)<br />

where =0corresponds to ground level, and where the -axis po<strong>in</strong>ts vertically upwards. F<strong>in</strong>d the extremal<br />

for the problem of m<strong>in</strong>imiz<strong>in</strong>g the total cost of the journey.<br />

The differential arc-length element of the flight path can be written as<br />

= p 2 + 2 = p 1+ 02 <br />

where 0 ≡ <br />

<br />

. Thus the cost <strong>in</strong>tegral to be m<strong>in</strong>imized is<br />

The function of this <strong>in</strong>tegral is<br />

=<br />

Z +<br />

−<br />

− =<br />

Z +<br />

−<br />

= −p 1+ 02<br />

The partial differentials required for the Euler equations are<br />

Therefore Euler’s equation equals<br />

−p 1+ 02 <br />

<br />

0 = 00 <br />

√ −<br />

− 02 <br />

√ −<br />

− 00 02 −<br />

1+<br />

02 1+<br />

02<br />

(1 + 02 ) 32<br />

<br />

= − −p 1+<br />

<br />

02<br />

<br />

− <br />

0 = −−p 1+ 02 − 00 <br />

√ −<br />

+ 02 <br />

√ −<br />

+ 00 02 −<br />

1+<br />

02 1+<br />

02<br />

(1 + 02 )<br />

This can be simplified by multiply<strong>in</strong>g the radical to give<br />

− − 2 02 − 04 − 00 − 00 02 + 02 + 04 + 00 02 =0<br />

Cancell<strong>in</strong>g terms gives<br />

00 + ¡ 1+ 02¢ =0<br />

Separat<strong>in</strong>g the variables leads to<br />

Z<br />

arctan 0 =<br />

0 Z<br />

02 +1 = − = − + 1<br />

32<br />

=0


5.4. SELECTION OF THE INDEPENDENT VARIABLE 113<br />

Integration gives<br />

() =<br />

Z <br />

−<br />

=<br />

Z <br />

−<br />

tan( 1 − ) = ln(cos( 1 − )) − ln(cos( 1 + ))<br />

<br />

ln<br />

+ 2 =<br />

³ ´<br />

cos(1−)<br />

cos( 1 +)<br />

+ 2<br />

<br />

Us<strong>in</strong>g the <strong>in</strong>itial condition that (−) =0gives 2 =0. Similarly the f<strong>in</strong>al condition () =0implies that<br />

1 =0. Thus Euler’s equation has determ<strong>in</strong>ed that the optimal trajectory that m<strong>in</strong>imizes the cost <strong>in</strong>tegral <br />

is<br />

() = 1 µ cos()<br />

ln cos()<br />

This example is typical of problems encountered <strong>in</strong> economics.<br />

5.4 Selection of the <strong>in</strong>dependent variable<br />

A wide selection of variables can be chosen as the <strong>in</strong>dependent variable for variational calculus. The derivation<br />

of Euler’s equation and example 51 both assumed that the <strong>in</strong>dependent variable is whereas example<br />

52 used as the <strong>in</strong>dependent variable, example 53 used , and Lagrange mechanics uses time as the<br />

<strong>in</strong>dependent variable. Selection of which variable to use as the <strong>in</strong>dependent variable does not change the<br />

physics of a problem, but some selections can simplify the mathematics for obta<strong>in</strong><strong>in</strong>g an analytic solution.<br />

The follow<strong>in</strong>g example of a cyl<strong>in</strong>drically-symmetric soap-bubble surface formed by blow<strong>in</strong>g a soap bubble that<br />

stretches between two circular hoops, illustrates the importance when select<strong>in</strong>g the <strong>in</strong>dependent variable.<br />

5.4 Example: Surface area of a cyl<strong>in</strong>drically-symmetric soap bubble<br />

Consider a cyl<strong>in</strong>drically-symmetric soap-bubble surface<br />

formed by blow<strong>in</strong>g a soap bubble that stretches between two<br />

circular hoops. The surface energy, that results from the surface<br />

tension of the soap bubble, is m<strong>in</strong>imized when the surface<br />

area of the bubble is m<strong>in</strong>imized. Assume that the axes of the<br />

two hoops lie along the axis as shown <strong>in</strong> the adjacent figure.<br />

It is <strong>in</strong>tuitively obvious that the soap bubble hav<strong>in</strong>g the m<strong>in</strong>imum<br />

surface area that is bounded by the two hoops will have<br />

a circular cross section that is concentric with the symmetry<br />

axis, and the radius will be smaller between the two hoops.<br />

Therefore, <strong>in</strong>tuition can be used to simplify the problem to<br />

f<strong>in</strong>d<strong>in</strong>g the shape of the contour of revolution around the axis<br />

of symmetry that def<strong>in</strong>es the shape of the surface of m<strong>in</strong>imum<br />

surface area. Use cyl<strong>in</strong>drical coord<strong>in</strong>ates ( ) and assume<br />

that hoop 1 at 1 has radius 1 and hoop 2 at 2 has radius<br />

2 . Consider the cases where either , or, areselectedto<br />

be the <strong>in</strong>dependent variable.<br />

The differential arc-length element of the circular annu-<br />

at constant between and + is given by =<br />

plus<br />

2 + 2 . Therefore the area of the <strong>in</strong>f<strong>in</strong>itessimal circular<br />

annulus is =2 which can be <strong>in</strong>tegrated to give the<br />

area of the surface ofthesoapbubbleboundedbythetwo<br />

circular hoops as<br />

Independent variable <br />

=2<br />

Z 2<br />

1<br />

p 2 + 2<br />

Assum<strong>in</strong>g that is the <strong>in</strong>dependent variable, then the surface area can be written as<br />

s<br />

Z 2<br />

µ 2 Z <br />

2<br />

=2 1+ =2 p 1+<br />

1 <br />

02 <br />

1<br />

z<br />

y<br />

z<br />

Cyl<strong>in</strong>drically-symmetric surface formed by<br />

rotation about the axis of a soap bubble<br />

suspended between two identical hoops<br />

centred on the axis.<br />

x


114 CHAPTER 5. CALCULUS OF VARIATIONS<br />

where 0 ≡ <br />

. The function of the surface <strong>in</strong>tegral is = p 1+ 02 The derivatives are<br />

and<br />

Therefore Euler’s equation gives<br />

This is not an easy equation to solve.<br />

Independent variable <br />

<br />

= p 1+ 02<br />

<br />

0 = <br />

q<br />

0<br />

1+( 0 ) 2<br />

⎛ ⎞<br />

<br />

⎝<br />

0<br />

q ⎠ − p 1+<br />

<br />

02 =0<br />

1+( 0 ) 2<br />

Consider the case where the <strong>in</strong>dependent variable is chosen to be , then the surface <strong>in</strong>tegral can be written<br />

as<br />

s<br />

Z 2<br />

µ 2 Z<br />

<br />

=2 1+ =2 p 1+<br />

1 <br />

02 <br />

where 0 ≡ <br />

. Thus the function of the surface <strong>in</strong>tegral is = √ 1+ 02 The derivatives are<br />

and<br />

Therefore Euler’s equation gives<br />

That is<br />

<br />

=0<br />

<br />

0 = <br />

q<br />

0<br />

1+( 0 ) 2<br />

⎛ ⎞<br />

0+ ⎝<br />

0<br />

q ⎠ =0<br />

<br />

1+( 0 ) 2<br />

0<br />

q1+( 0 ) 2 = <br />

where is a constant. This can be rewritten as<br />

02 ¡ 2 − 2¢ = 2<br />

or<br />

0 = <br />

= <br />

p<br />

2 − 2<br />

The <strong>in</strong>tegral of this is<br />

³ = cosh −1 <br />

´<br />

+ <br />

<br />

That is<br />

= cosh − <br />

<br />

which is the equation of a catenary. The catenary is the shape of a uniform flexible cable hung <strong>in</strong> a uniform<br />

gravitational field. The constants and are given by the end po<strong>in</strong>ts. The physics of the solution must be<br />

identical for either choice of <strong>in</strong>dependent variable. However, mathematically one case is easier to solve than<br />

the other because, <strong>in</strong> the latter case, one term <strong>in</strong> Euler’s equation is zero.


5.5. FUNCTIONS WITH SEVERAL INDEPENDENT VARIABLES () 115<br />

5.5 Functions with several <strong>in</strong>dependent variables ()<br />

The discussion has focussed on systems hav<strong>in</strong>g only a s<strong>in</strong>gle function () such that the functional is an<br />

extremum. It is more common to have a functional that is dependent upon several <strong>in</strong>dependent variables<br />

[ 1 ()1 0 () 2()2 0 () ; ] which can be written as<br />

=<br />

Z 2<br />

1<br />

X<br />

=1<br />

[ () 0 (); ] (5.16)<br />

where =1 2 3 <br />

By analogy with the one dimensional problem, def<strong>in</strong>e neighbor<strong>in</strong>g functions for each variable. Then<br />

( ) = (0)+ () (5.17)<br />

( 0 ) ≡ ( )<br />

= (0)<br />

+ <br />

<br />

where are <strong>in</strong>dependent functions of that vanish at 1 and 2 Us<strong>in</strong>g equations 512 and 517 leads to<br />

the requirements for an extremum value to be<br />

<br />

= Z 2<br />

1<br />

X<br />

<br />

µ <br />

<br />

<br />

+ <br />

0 <br />

<br />

0 Z 2<br />

=<br />

<br />

1<br />

X<br />

<br />

µ <br />

− <br />

<br />

<br />

0 <br />

<br />

() =0 (5.18)<br />

If the variables () are <strong>in</strong>dependent, thenthe () are <strong>in</strong>dependent. S<strong>in</strong>ce the () are <strong>in</strong>dependent,<br />

then evaluat<strong>in</strong>g the above equation at =0implies that each term <strong>in</strong> the bracket must vanish <strong>in</strong>dependently.<br />

That is, Euler’s differential equation becomes a set of equations for the <strong>in</strong>dependent variables<br />

<br />

− <br />

<br />

0 =0 (5.19)<br />

where =1 2 3 Thus, each of the equations can be solved <strong>in</strong>dependently when the variables are<br />

<strong>in</strong>dependent. Note that Euler’s equation <strong>in</strong>volves partial derivatives for the dependent variables , 0 and<br />

the total derivative for the <strong>in</strong>dependent variable .<br />

5.5 Example: Fermat’s Pr<strong>in</strong>ciple<br />

In 1662 Fermat’s proposed that the propagation of<br />

light obeyed the generalized pr<strong>in</strong>ciple of least transit time.<br />

In optics, Fermat’s pr<strong>in</strong>ciple, or the pr<strong>in</strong>ciple of least<br />

P 1<br />

(0, y , 0)<br />

1<br />

1<br />

time, is the pr<strong>in</strong>ciple that the path taken between two<br />

po<strong>in</strong>ts by a ray of light is the path that can be traversed <strong>in</strong><br />

the least time. Historically, the proof of Fermat’s pr<strong>in</strong>ciple<br />

(x, 0, z)<br />

by Johann Bernoulli was one of the first triumphs of<br />

O<br />

the calculus of variations, and served as a guid<strong>in</strong>g pr<strong>in</strong>ciple<br />

<strong>in</strong> the formulation of physical laws us<strong>in</strong>g variational<br />

calculus.<br />

2<br />

Consider the geometry shown <strong>in</strong> the figure, where<br />

the light travels from the po<strong>in</strong>t 1 (0 1 0) to the po<strong>in</strong>t<br />

P 2<br />

2 ( 2 − 2 0). The light beam <strong>in</strong>tersects a plane glass<br />

Light <strong>in</strong>cident upon a plane glass <strong>in</strong>terface <strong>in</strong> the<br />

<strong>in</strong>terface at the po<strong>in</strong>t ( 0).<br />

( ) plane at =0.<br />

The French mathematician Fermat discovered that<br />

the required path travelled by light is the path for which<br />

the travel time is a m<strong>in</strong>imum. That is, the transit time from the <strong>in</strong>itial po<strong>in</strong>t 1 to the f<strong>in</strong>al po<strong>in</strong>t 2 is<br />

given by<br />

=<br />

Z 2<br />

1<br />

=<br />

Z 2<br />

1<br />

<br />

= 1 <br />

Z 2<br />

1<br />

= 1 <br />

Z 2<br />

1<br />

q<br />

( ) 1+( 0 ) 2 +( 0 ) 2 <br />

(x , -y , 0)<br />

2 2<br />

assum<strong>in</strong>g that the velocity of light <strong>in</strong> any medium is given by = where is the refractive <strong>in</strong>dex of the<br />

medium and is the velocity of light <strong>in</strong> vacuum.


116 CHAPTER 5. CALCULUS OF VARIATIONS<br />

This is a problem that has two dependent variables () and () with chosen as the <strong>in</strong>dependent<br />

variable. The <strong>in</strong>tegral can be broken <strong>in</strong>to two parts 1 → 0 and 0 →− 2 <br />

= 1 ∙Z 0 q<br />

1 1+(<br />

<br />

0 ) 2 +( 0 ) 2 +<br />

1<br />

Z −2<br />

0<br />

¸<br />

2<br />

q1+( 0 ) 2 +( 0 ) 2 <br />

The functionals are functions of 0 and 0 but not or . Thus Euler’s equation for simplifies to<br />

0+ µ 1<br />

( 1 0<br />

√<br />

1+ 0 2<br />

+ + 2 0 <br />

√ 02 1+<br />

02<br />

+ ) =0<br />

0 2<br />

This implies that 0 =0,therefore is a constant. S<strong>in</strong>ce the <strong>in</strong>itial and f<strong>in</strong>al values were chosen to be<br />

1 = 2 =0, therefore at the <strong>in</strong>terface =0. Similarly Euler’s equations for are<br />

0+ µ 1<br />

( 1 0<br />

√<br />

1+ 0 2<br />

+ + 2 0 <br />

√ 02 1+<br />

02<br />

+ ) =0<br />

0 2<br />

But 0 =tan 1 for 1 and 0 = − tan 2 for 2 and it was shown that 0 =0. Thus<br />

⎛<br />

⎞<br />

0+ ⎝ 1<br />

( 1 tan 1 2 tan 2<br />

−<br />

) ⎠ = q1+(tan 1 )<br />

q1+(tan µ 1 2<br />

2 ) 2 ( 1 s<strong>in</strong> 1 − 2 s<strong>in</strong> 2 )<br />

=0<br />

Therefore 1 ( 1 s<strong>in</strong> 1 − 2 s<strong>in</strong> 2 )=constant which must be zero s<strong>in</strong>ce when 1 = 2 then 1 = 2 . Thus<br />

Fermat’s pr<strong>in</strong>ciple leads to Snell’s Law.<br />

1 s<strong>in</strong> 1 = 2 s<strong>in</strong> 2<br />

The geometry of this problem is simple enough to directly m<strong>in</strong>imize the path rather than us<strong>in</strong>g Euler’s<br />

equations for the two parameters as performed above. The lengths of the paths 1 and 2 are<br />

q<br />

1 = 2 + 1 2 + 2<br />

q<br />

2 = ( 2 − ) 2 + 2 2 + 2<br />

The total transit time is given by<br />

= 1 <br />

µ 1<br />

q<br />

<br />

2 + 1 2 + 2 + 2<br />

q( 2 − ) 2 + 2 2 + 2<br />

This problem <strong>in</strong>volves two dependent variables, () and (). Tof<strong>in</strong>d the m<strong>in</strong>ima, set the partial derivatives<br />

<br />

<br />

<br />

=0and<br />

=0. That is,<br />

<br />

= 1 ( 1 <br />

p<br />

2 + 1 2 + + 2 <br />

)=0<br />

q( 2 2 − ) 2 + 2 2 + 2<br />

This is zero only if<br />

=0,thatisthepo<strong>in</strong>t lies <strong>in</strong> the plane conta<strong>in</strong><strong>in</strong>g 1 and 2 . Similarly<br />

<br />

= 1 ( 1 <br />

p<br />

2 + 1 2 + − 2 ( 2 − )<br />

)= q( 1 2 2 − ) 2 + 2 2 + ( 1 s<strong>in</strong> 1 − 2 s<strong>in</strong> 2 )=0<br />

2<br />

This is zero only if Snell’s law applies that is<br />

1 s<strong>in</strong> 1 = 2 s<strong>in</strong> 2<br />

Fermat’s pr<strong>in</strong>ciple has shown that the refracted light is given by Snell’s Law, and is <strong>in</strong> a plane normal to the<br />

surface. The laws of reflection also are given s<strong>in</strong>ce then 1 = 2 = and the angle of reflection equals the<br />

angle of <strong>in</strong>cidence.


5.6. EULER’S INTEGRAL EQUATION 117<br />

5.6 Example: M<strong>in</strong>imum of (∇) 2 <strong>in</strong> a volume<br />

F<strong>in</strong>d the function ( 1 2 3 ) that has the m<strong>in</strong>imum value of (∇) 2 perunitvolume. Forthevolume<br />

it is desired to m<strong>in</strong>imize the follow<strong>in</strong>g<br />

= 1 Z Z Z<br />

(∇) 2 1 2 3 = 1 Z Z Z " µ 2 µ 2 µ # 2 <br />

+ + 1 2 3<br />

<br />

<br />

1 2 3<br />

Note that the variables 1 2 3 are <strong>in</strong>dependent, and thus Euler’s equation for several <strong>in</strong>dependent variables<br />

can be used. To m<strong>in</strong>imize the functional , the function<br />

µ 2 µ 2 µ 2<br />

<br />

= + +<br />

()<br />

1 2 3<br />

must satisfy the Euler equation<br />

<br />

3X<br />

µ <br />

− <br />

<br />

=1 0 =0<br />

where 0 = <br />

<br />

. Substitute <strong>in</strong>to Euler’s equation gives<br />

3X<br />

=1<br />

µ <br />

<br />

=0<br />

<br />

This is just Laplace’s equation<br />

∇ 2 =0<br />

Therefore must satisfy Laplace’s equation <strong>in</strong> order that the functional be a m<strong>in</strong>imum.<br />

5.6 Euler’s <strong>in</strong>tegral equation<br />

An <strong>in</strong>tegral form of the Euler differential equation can be written which is useful for cases when the function<br />

does not depend explicitly on the <strong>in</strong>dependent variable , thatis,when <br />

<br />

=0 Note that<br />

<br />

= <br />

+ <br />

+ 0<br />

0 (5.20)<br />

<br />

But<br />

µ<br />

<br />

0 <br />

0 = 0<br />

0 + <br />

0<br />

0 (5.21)<br />

Comb<strong>in</strong><strong>in</strong>g these two equations gives<br />

µ<br />

<br />

0 <br />

0 = <br />

− <br />

− 0<br />

+ <br />

0<br />

0 (5.22)<br />

The last two terms can be rewritten as µ <br />

0 <br />

0 − <br />

(5.23)<br />

<br />

which vanishes when the Euler equation is satisfied. Therefore the above equation simplifies to<br />

<br />

− <br />

<br />

µ<br />

− 0 <br />

0 <br />

=0 (5.24)<br />

This <strong>in</strong>tegral form of Euler’s equation is especially useful when <br />

<br />

=0 that is, when does not depend<br />

explicitly on the <strong>in</strong>dependent variable . Then the first <strong>in</strong>tegral of equation 524 is a constant, i.e.<br />

− 0 = constant (5.25)<br />

0 This is Euler’s <strong>in</strong>tegral variational equation. Note that the shortest distance between two po<strong>in</strong>ts, the m<strong>in</strong>imum<br />

surface of rotation, and the brachistochrone, described earlier, all are examples where <br />

<br />

=0and thus<br />

the <strong>in</strong>tegral form of Euler’s equation is useful for solv<strong>in</strong>g these cases.


118 CHAPTER 5. CALCULUS OF VARIATIONS<br />

5.7 Constra<strong>in</strong>ed variational systems<br />

Impos<strong>in</strong>g a constra<strong>in</strong>t on a variational system implies:<br />

1. The constra<strong>in</strong>ed coord<strong>in</strong>ates () are correlated which violates<br />

the assumption made <strong>in</strong> chapter 55 that the variables are <strong>in</strong>dependent.<br />

F f<br />

y<br />

N<br />

2. Constra<strong>in</strong>ed motion implies that constra<strong>in</strong>t forces must be act<strong>in</strong>g<br />

to account for the correlation of the variables. These constra<strong>in</strong>t<br />

forces must be taken <strong>in</strong>to account <strong>in</strong> the equations of motion.<br />

mg<br />

For example, for a disk roll<strong>in</strong>g down an <strong>in</strong>cl<strong>in</strong>ed plane without slipp<strong>in</strong>g,<br />

there are three coord<strong>in</strong>ates [perpendicular to the wedge], ,[Along<br />

the surface of the wedge], and the rotation angle shown <strong>in</strong> figure 52<br />

The constra<strong>in</strong>t forces, F N, lead to the correlation of the variables such<br />

that = , while = . Basically there is only one <strong>in</strong>dependent<br />

variable, which can be either or The use of only one <strong>in</strong>dependent<br />

variable essentially buries the constra<strong>in</strong>t forces under the rug, which is<br />

Figure 5.2: A disk roll<strong>in</strong>g down<br />

an <strong>in</strong>cl<strong>in</strong>ed plane.<br />

f<strong>in</strong>e if you only need to know the equation of motion. If you need to determ<strong>in</strong>e the forces of constra<strong>in</strong>t then<br />

it is necessary to <strong>in</strong>clude all coord<strong>in</strong>ates explicitly <strong>in</strong> the equations of motion as discussed below.<br />

5.7.1 Holonomic constra<strong>in</strong>ts<br />

Most systems <strong>in</strong>volve restrictions or constra<strong>in</strong>ts that couple the coord<strong>in</strong>ates. For example, the () may<br />

be conf<strong>in</strong>ed to a surface <strong>in</strong> coord<strong>in</strong>ate space. The constra<strong>in</strong>ts mean that the coord<strong>in</strong>ates () are not <strong>in</strong>dependent,<br />

but are related by equations of constra<strong>in</strong>t. A constra<strong>in</strong>t is called holonomic if the equations of<br />

constra<strong>in</strong>t can be expressed <strong>in</strong> the form of an algebraic equation that directly and unambiguously specifies<br />

the shape of the surface of constra<strong>in</strong>t. A non-holonomic constra<strong>in</strong>t does not provide an algebraic relation<br />

between the correlated coord<strong>in</strong>ates. In addition to the holonomy of the constra<strong>in</strong>ts, the equations of constra<strong>in</strong>t<br />

also can be grouped <strong>in</strong>to the follow<strong>in</strong>g three classifications depend<strong>in</strong>g on whether they are algebraic,<br />

differential, or <strong>in</strong>tegral. These three classifications for the constra<strong>in</strong>ts exhibit different holonomy relat<strong>in</strong>g the<br />

coupled coord<strong>in</strong>ates. Fortunately the solution of constra<strong>in</strong>ed systems is greatly simplified if the equations of<br />

constra<strong>in</strong>t are holonomic.<br />

5.7.2 Geometric (algebraic) equations of constra<strong>in</strong>t<br />

Geometric constra<strong>in</strong>ts can be expressed <strong>in</strong> the form of algebraic relations that directly specify the shape of<br />

the surface of constra<strong>in</strong>t <strong>in</strong> coord<strong>in</strong>ate space 1 2 <br />

( 1 2 ; ) =0 (5.26)<br />

where =1 2 3 . Therecanbe such equations of constra<strong>in</strong>t where 0 ≤ ≤ . Anexampleofsucha<br />

geometric constra<strong>in</strong>t is when the motion is conf<strong>in</strong>ed to the surface of a sphere of radius <strong>in</strong> coord<strong>in</strong>ate space<br />

whichcanbewritten<strong>in</strong>theform = 2 + 2 + 2 − 2 =0 Such algebraic constra<strong>in</strong>t equations are called<br />

Holonomic which allows use of generalized coord<strong>in</strong>ates as well as Lagrange multipliers to handle both the<br />

constra<strong>in</strong>t forces and the correlation of the coord<strong>in</strong>ates.<br />

5.7.3 K<strong>in</strong>ematic (differential) equations of constra<strong>in</strong>t<br />

The constra<strong>in</strong>t equations also can be expressed <strong>in</strong> terms of the <strong>in</strong>f<strong>in</strong>itessimal displacements of the form<br />

X <br />

+ <br />

=0 (5.27)<br />

<br />

=1<br />

where =1 2 3 , =1 2 3. If equation (527) represents the total differential of a function then<br />

it can be <strong>in</strong>tegrated to give a holonomic relation of the form of equation 526. However, if equation 527 is


5.7. CONSTRAINED VARIATIONAL SYSTEMS 119<br />

not the total differential, then it is non-holonomic and can be <strong>in</strong>tegrated only after hav<strong>in</strong>g solved the full<br />

problem.<br />

An example of differential constra<strong>in</strong>t equations is for a wheel roll<strong>in</strong>g on a plane without slipp<strong>in</strong>g which is<br />

non-holonomic and more complicated than might be expected. The wheel mov<strong>in</strong>g on a plane has five degrees<br />

of freedom s<strong>in</strong>ce the height is fixed. That is, the motion of the center of mass requires two coord<strong>in</strong>ates<br />

( ) plus there are three angles ( ) where is the rotation angle for the wheel, is the pivot angle of<br />

the axis, and is the tilt angle of the wheel. If the wheel slides then all five degrees of freedom are active.<br />

If the axis of rotation of the wheel is horizontal, that is, the tilt angle =0is constant, then this k<strong>in</strong>ematic<br />

system leads to three differential constra<strong>in</strong>t equations The wheel can roll with angular velocity ˙, aswellas<br />

pivot which corresponds to a change <strong>in</strong> Comb<strong>in</strong><strong>in</strong>g these leads to two differential equations of constra<strong>in</strong>t<br />

− s<strong>in</strong> =0 + cos =0 (5.28)<br />

These constra<strong>in</strong>ts are <strong>in</strong>sufficient to provide f<strong>in</strong>ite relations between all the coord<strong>in</strong>ates. That is, the constra<strong>in</strong>ts<br />

cannot be reduced by <strong>in</strong>tegration to the form of equation 526 because there is no functional relation<br />

between and the other three variables, . Many roll<strong>in</strong>g trajectories are possible between any two po<strong>in</strong>ts<br />

of contact on the plane that are related to different pivot angles. That is, the po<strong>in</strong>t of contact of the disk<br />

could pivot plus roll <strong>in</strong> a circle return<strong>in</strong>g to the same po<strong>in</strong>t where are unchanged whereas the value<br />

of depends on the circumference of the circle. As a consequence the roll<strong>in</strong>g constra<strong>in</strong>t is non-holonomic<br />

except for the case where the disk rolls <strong>in</strong> a straight l<strong>in</strong>e and rema<strong>in</strong>s vertical.<br />

5.7.4 Isoperimetric (<strong>in</strong>tegral) equations of constra<strong>in</strong>t<br />

Equations of constra<strong>in</strong>t also can be expressed <strong>in</strong> terms of direct <strong>in</strong>tegrals. This situation is encountered for<br />

isoperimetric problems, such as f<strong>in</strong>d<strong>in</strong>g the maximum volume bounded by a surface of fixed area, or the<br />

shape of a hang<strong>in</strong>g rope of fixed length. Integral constra<strong>in</strong>ts occur <strong>in</strong> economics when m<strong>in</strong>imiz<strong>in</strong>g some cost<br />

algorithm subject to a fixed total cost constra<strong>in</strong>t.<br />

A simple example of an isoperimetric problem <strong>in</strong>volves f<strong>in</strong>d<strong>in</strong>g the curve = () such that the functional<br />

has an extremum where the curve () satisfies boundary conditions such that ( 1 )= and ( 2 )=,<br />

that is<br />

Z 2<br />

() = ( 0 ; ) (5.29)<br />

1<br />

is an extremum such that the perimeter also is constra<strong>in</strong>ed to satisfy<br />

Z 2<br />

() = ( 0 ; ) = (5.30)<br />

1<br />

where is a fixed length. This <strong>in</strong>tegral constra<strong>in</strong>t is geometric and holonomic. Another example is f<strong>in</strong>d<strong>in</strong>g<br />

them<strong>in</strong>imumsurfaceareaofaclosedsurfacesubjecttotheenclosedvolumebe<strong>in</strong>gtheconstra<strong>in</strong>t.<br />

5.7.5 Properties of the constra<strong>in</strong>t equations<br />

Holonomic constra<strong>in</strong>ts Geometric constra<strong>in</strong>ts can be expressed <strong>in</strong> the form of an algebraic equation<br />

that directly specifies the shape of the surface of constra<strong>in</strong>t<br />

( 1 2 3 ; ) =0 (5.31)<br />

Such a system is called holonomic s<strong>in</strong>ce there is a direct relation between the coupled variables. An example<br />

of such a holonomic geometric constra<strong>in</strong>t is if the motion is conf<strong>in</strong>ed to the surface of a sphere of radius <br />

which can be written <strong>in</strong> the form<br />

= 2 + 2 + 2 − 2 =0 (5.32)<br />

Non-holonomic constra<strong>in</strong>ts There are many classifications of non-holonomic constra<strong>in</strong>ts that exist<br />

if equation (531) is not satisfied. The algebraic approach is difficult to handle when the constra<strong>in</strong>t is an<br />

<strong>in</strong>equality, such as the requirement that the location is restricted to lie <strong>in</strong>side a spherical shell of radius <br />

which can be expressed as<br />

= 2 + 2 + 2 − 2 ≤ 0 (5.33)


120 CHAPTER 5. CALCULUS OF VARIATIONS<br />

This non-holonomic constra<strong>in</strong>ed system has a one-sided constra<strong>in</strong>t. Systems usually are non-holonomic if<br />

the constra<strong>in</strong>t is k<strong>in</strong>ematic as discussed above.<br />

Partial Holonomic constra<strong>in</strong>ts Partial-holonomic constra<strong>in</strong>ts are holonomic for a restricted range<br />

of the constra<strong>in</strong>t surface <strong>in</strong> coord<strong>in</strong>ate space, and this range can be case specific. This can occur if the<br />

constra<strong>in</strong>t force is one-sided and perpendicular to the path. An example is the pendulum with the mass<br />

attached to the fulcrum by a flexible str<strong>in</strong>g that provides tension but not compression. Then the pendulum<br />

length is constant only if the tension <strong>in</strong> the str<strong>in</strong>g is positive. Thus the pendulum will be holonomic if<br />

the gravitational plus centrifugal forces are such that the tension <strong>in</strong> the str<strong>in</strong>g is positive, but the system<br />

becomes non-hononomic if the tension is negative as can happen when the pendulum rotates to an upright<br />

angle where the centrifugal force outwards is <strong>in</strong>sufficient to compensate for the vertical downward component<br />

of the gravitational force. There are many other examples where the motion of an object is holonomic when<br />

the object is pressed aga<strong>in</strong>st the constra<strong>in</strong>t surface, such as the surface of the Earth, but is unconstra<strong>in</strong>ed if<br />

the object leaves the surface.<br />

Time dependence<br />

A constra<strong>in</strong>t is called scleronomic if the constra<strong>in</strong>t is not explicitly time dependent. This ignores the time<br />

dependence conta<strong>in</strong>ed with<strong>in</strong> the solution of the equations of motion. Fortunately a major fraction of<br />

systems are scleronomic. The constra<strong>in</strong>t is called rheonomic if the constra<strong>in</strong>t is explicitly time dependent.<br />

Anexampleofarheonomicsystemiswherethesizeorshape of the surface of constra<strong>in</strong>t is explicitly time<br />

dependent such as a deflat<strong>in</strong>g pneumatic tire.<br />

Energy conservation<br />

The solution depends on whether the constra<strong>in</strong>t is conservative or dissipative, that is, if friction or drag are<br />

act<strong>in</strong>g. The system will be conservative if there are no drag forces, and the constra<strong>in</strong>t forces are perpendicular<br />

to the trajectory of the path such as the motion of a charged particle <strong>in</strong> a magnetic field. Forces of constra<strong>in</strong>t<br />

can result from slid<strong>in</strong>g of two solid surfaces, roll<strong>in</strong>g of solid objects, fluid flow <strong>in</strong> a liquid or gas, or result from<br />

electromagnetic forces. Energy dissipation can result from friction, drag <strong>in</strong> a fluid or gas, or f<strong>in</strong>ite resistance<br />

of electric conductors lead<strong>in</strong>g to dissipation of <strong>in</strong>duced electric currents <strong>in</strong> a conductor, e.g. eddy currents.<br />

A roll<strong>in</strong>g constra<strong>in</strong>t is unusual <strong>in</strong> that friction between the roll<strong>in</strong>g bodies is necessary to ma<strong>in</strong>ta<strong>in</strong> roll<strong>in</strong>g.<br />

A disk on a frictionless <strong>in</strong>cl<strong>in</strong>ed plane will conserve it’s angular momentum s<strong>in</strong>ce there is no torque act<strong>in</strong>g<br />

if the roll<strong>in</strong>g contact is frictionless, that is, the disk will just slide. If the friction is sufficient to stop slid<strong>in</strong>g,<br />

then the bodies will roll and not slide. A perfect roll<strong>in</strong>g body does not dissipate energy s<strong>in</strong>ce no work is<br />

done at the <strong>in</strong>stantaneous po<strong>in</strong>t of contact where both bodies are <strong>in</strong> zero relative motion and the force is<br />

perpendicular to the motion. In real life, a roll<strong>in</strong>g wheel can <strong>in</strong>volve a very small energy dissipation due to<br />

deformation at the po<strong>in</strong>t of contact coupled with non-elastic properties of the material used to make the<br />

wheel and the plane surface. For example, a pneumatic tire can heat up and expand due to flex<strong>in</strong>g of the<br />

tire.<br />

5.7.6 Treatment of constra<strong>in</strong>t forces <strong>in</strong> variational calculus<br />

There are three major approaches to handle constra<strong>in</strong>t forces <strong>in</strong> variational calculus. All three of them exploit<br />

the tremendous freedom and flexibility available when us<strong>in</strong>g generalized coord<strong>in</strong>ates. The (1) generalized<br />

coord<strong>in</strong>ate approach, described <strong>in</strong> chapter 58, exploits the correlation of the coord<strong>in</strong>ates due to the <br />

constra<strong>in</strong>t forces to reduce the dimension of the equations of motion to = − degrees of freedom. This<br />

approach embeds the constra<strong>in</strong>t forces, <strong>in</strong>to the choice of generalized coord<strong>in</strong>ates and does not determ<strong>in</strong>e<br />

the constra<strong>in</strong>t forces, (2) Lagrange multiplier approach, described <strong>in</strong> chapter 59, exploits generalized<br />

coord<strong>in</strong>ates but <strong>in</strong>cludes the constra<strong>in</strong>t forces <strong>in</strong>to the Euler equations to determ<strong>in</strong>e both the constra<strong>in</strong>t<br />

forces <strong>in</strong> addition to the equations of motion. (3) Generalized forces approach, described <strong>in</strong> chapter<br />

673 <strong>in</strong>troduces constra<strong>in</strong>t and other forces explicitly.


5.8. GENERALIZED COORDINATES IN VARIATIONAL CALCULUS 121<br />

5.8 Generalized coord<strong>in</strong>ates <strong>in</strong> variational calculus<br />

Newtonian mechanics is based on a vectorial treatment of mechanics which can be difficult to apply when<br />

solv<strong>in</strong>g complicated problems <strong>in</strong> mechanics. Constra<strong>in</strong>t forces act<strong>in</strong>g on a system usually are unknown. In<br />

Newtonian mechanics constra<strong>in</strong>ed forces must be <strong>in</strong>cluded explicitly so that they can be determ<strong>in</strong>ed simultaneously<br />

with the solution of the dynamical equations of motion. The major advantage of the variational<br />

approaches is that solution of the dynamical equations of motion can be simplified by express<strong>in</strong>g the motion<br />

<strong>in</strong> terms of <strong>in</strong>dependent generalized coord<strong>in</strong>ates. These generalized coord<strong>in</strong>ates can be any set of <strong>in</strong>dependent<br />

variables, ,where1 ≤ ≤ , plus the correspond<strong>in</strong>g velocities ˙ for Lagrangian mechanics,<br />

or the correspond<strong>in</strong>g canonical variables, for Hamiltonian mechanics. These generalized coord<strong>in</strong>ates for<br />

the variables are used to specify the scalar functional dependence on these generalized coord<strong>in</strong>ates. The<br />

variational approach employs this scalar functional to determ<strong>in</strong>e the trajectory. The generalized coord<strong>in</strong>ates<br />

used for the variational approach do not need to be orthogonal, they only need to be <strong>in</strong>dependent s<strong>in</strong>ce they<br />

are used only to completely specify the magnitude of the scalar functional. This greatly expands the arsenal<br />

of possible generalized coord<strong>in</strong>ates beyond what is available us<strong>in</strong>g Newtonian mechanics. For example,<br />

generalized coord<strong>in</strong>ates can be the dimensionless amplitudes for the normal modes of coupled oscillator<br />

systems, or action-angle variables. In addition, generalized coord<strong>in</strong>ates hav<strong>in</strong>g different dimensions can be<br />

used for each of the variables. Each generalized coord<strong>in</strong>ate, specifies an <strong>in</strong>dependent mode of the system,<br />

not a specific particle. For example, each normal mode of coupled oscillators can <strong>in</strong>volve correlated motion of<br />

several coupled particles. The major advantage of us<strong>in</strong>g generalized coord<strong>in</strong>ates is that they can be chosen<br />

to be perpendicular to a correspond<strong>in</strong>g constra<strong>in</strong>t force, and therefore that specific constra<strong>in</strong>t force does no<br />

work for motion along that generalized coord<strong>in</strong>ate. Moreover, the constra<strong>in</strong>ed motion does no work <strong>in</strong> the<br />

direction of the constra<strong>in</strong>t force for rigid constra<strong>in</strong>ts. Thus generalized coord<strong>in</strong>ates allow specific constra<strong>in</strong>t<br />

forces to be ignored <strong>in</strong> evaluation of the m<strong>in</strong>imized functional. This freedom and flexibility of choice of generalized<br />

coord<strong>in</strong>ates allows the correlated motion produced by the constra<strong>in</strong>t forces to be embedded directly<br />

<strong>in</strong>to the choice of the <strong>in</strong>dependent generalized coord<strong>in</strong>ates, and the actual constra<strong>in</strong>t forces can be ignored.<br />

Embedd<strong>in</strong>g of the constra<strong>in</strong>t <strong>in</strong>duced correlations <strong>in</strong>to the generalized coord<strong>in</strong>ates, effectively “sweeps the<br />

constra<strong>in</strong>t forces under the rug” which greatly simplifies the equations of motion for any system that <strong>in</strong>volve<br />

constra<strong>in</strong>t forces. Selection of the appropriate generalized coord<strong>in</strong>ates can be obvious, and often it is<br />

performed subconsciously by the user.<br />

Three variational approaches are used that employ generalized coord<strong>in</strong>ates to derive the equations of<br />

motion of a system that has generalized coord<strong>in</strong>ates subject to constra<strong>in</strong>ts.<br />

1) M<strong>in</strong>imal set of generalized coord<strong>in</strong>ates: When the equations of constra<strong>in</strong>t are holonomic, then<br />

the algebraic constra<strong>in</strong>t relations can be used to transform the coord<strong>in</strong>ates <strong>in</strong>to = − <strong>in</strong>dependent<br />

generalized coord<strong>in</strong>ates . This approach reduces the number of unknowns, by the number of constra<strong>in</strong>ts<br />

, to give a m<strong>in</strong>imal set of = − <strong>in</strong>dependent generalized dynamical variables. The forces of constra<strong>in</strong>t<br />

are not explicitly discussed, or determ<strong>in</strong>ed, when this generalized coord<strong>in</strong>ate approach is employed. This<br />

approach greatly simplifies solution of dynamical problems by avoid<strong>in</strong>g the need for explicit treatment of the<br />

constra<strong>in</strong>t forces. This approach is straight forward for holonomic constra<strong>in</strong>ts, s<strong>in</strong>ce the spatial coord<strong>in</strong>ates<br />

1 () () are coupled by algebraic equations which can be used to make the transformation to<br />

generalized coord<strong>in</strong>ates. Thus the coupled spatial coord<strong>in</strong>ates are transformed to = − <strong>in</strong>dependent<br />

generalized dynamical coord<strong>in</strong>ates 1 () (), and their generalized first derivatives ˙ 1 () ˙ () These<br />

generalized coord<strong>in</strong>ates are <strong>in</strong>dependent, and thus it is possible to use Euler’s equation for each <strong>in</strong>dependent<br />

parameter <br />

<br />

− <br />

=0 (5.34)<br />

<br />

where =1 2 3 There are = − such Euler equations. The freedom to choose generalized coord<strong>in</strong>ates<br />

underlies the tremendous advantage of apply<strong>in</strong>g the variational approach.<br />

2) Lagrange multipliers: The Lagrange equations, plus the equations of constra<strong>in</strong>t, can be used<br />

to explicitly determ<strong>in</strong>e the generalized coord<strong>in</strong>ates plus the constra<strong>in</strong>t forces. That is, + unknowns<br />

are determ<strong>in</strong>ed. This approach is discussed <strong>in</strong> chapter 59.<br />

3) Generalized forces: This approach <strong>in</strong>troduces the constra<strong>in</strong>t forces explicity. This approach, applied<br />

to Lagrangian mechanics, is discussed <strong>in</strong> chapter 663<br />

The above three approaches exploit generalized coord<strong>in</strong>ates to handle constra<strong>in</strong>t forces as described <strong>in</strong><br />

chapter 6<br />

0


122 CHAPTER 5. CALCULUS OF VARIATIONS<br />

5.9 Lagrange multipliers for holonomic constra<strong>in</strong>ts<br />

5.9.1 Algebraic equations of constra<strong>in</strong>t<br />

The Lagrange multiplier technique provides a powerful, and elegant, way to handle holonomic constra<strong>in</strong>ts<br />

us<strong>in</strong>g Euler’s equations 1 . The general method of Lagrange multipliers for variables, with constra<strong>in</strong>ts,<br />

is best <strong>in</strong>troduced us<strong>in</strong>g Bernoulli’s <strong>in</strong>genious exploitation of virtual <strong>in</strong>f<strong>in</strong>itessimal displacements, which<br />

Lagrange signified by the symbol . The term “virtual” refers to an <strong>in</strong>tentional variation of the generalized<br />

coord<strong>in</strong>ates <strong>in</strong> order to elucidate the local sensitivity of a function ( ) to variation of the variable.<br />

Contrary to the usual <strong>in</strong>f<strong>in</strong>itessimal <strong>in</strong>terval <strong>in</strong> differential calculus, where an actual displacement occurs<br />

dur<strong>in</strong>g a time , a virtual displacement is imag<strong>in</strong>ed to be an <strong>in</strong>stantaneous, <strong>in</strong>f<strong>in</strong>itessimal, displacement of<br />

a coord<strong>in</strong>ate, not an actual displacement, <strong>in</strong> order to elucidate the local dependence of on the coord<strong>in</strong>ate.<br />

The local dependence of any functional to virtual displacements of all coord<strong>in</strong>ates, is given by tak<strong>in</strong>g<br />

the partial differentials of .<br />

X <br />

= (5.35)<br />

<br />

<br />

The function is stationary, that is an extremum, if equation 535 equals zero. The extremum of the<br />

functional , given by equation 516 can be expressed <strong>in</strong> a compact form us<strong>in</strong>g the virtual displacement<br />

formalism as<br />

Z 2<br />

= <br />

1<br />

X<br />

<br />

[ () 0 (); ] =<br />

X<br />

<br />

<br />

<br />

=0 (5.36)<br />

The auxiliary conditions, due to the holonomic algebraic constra<strong>in</strong>ts for the variables ,canbe<br />

expressed by the equations<br />

(q) =0 (5.37)<br />

where 1 ≤ ≤ and 1 ≤ ≤ with . The variational problem for the holonomic constra<strong>in</strong>t<br />

equations also can be written <strong>in</strong> terms of differential equations where 1 ≤ ≤ <br />

=<br />

X<br />

=1<br />

<br />

<br />

=0 (5.38)<br />

S<strong>in</strong>ce equations 536 and 538 both equal zero, the equations 538 can be multiplied by arbitrary<br />

undeterm<strong>in</strong>ed factors and added to equations 536 to give.<br />

( )+ 1 1 + 2 2 ·· ·· =0 (5.39)<br />

Note that this is not trivial <strong>in</strong> that although the sum of the constra<strong>in</strong>t equations for each is zero; the<br />

<strong>in</strong>dividual terms of the sum are not zero.<br />

Insert equations 536 plus 538 <strong>in</strong>to 539 and collect all terms, gives<br />

X<br />

<br />

Ã<br />

<br />

<br />

+<br />

X<br />

=1<br />

!<br />

<br />

=0 (5.40)<br />

<br />

Note that all the are free <strong>in</strong>dependent variations and thus the terms <strong>in</strong> the brackets, which are the<br />

coefficients of each , <strong>in</strong>dividually must equal zero. For each of the values of , the correspond<strong>in</strong>g bracket<br />

implies<br />

<br />

<br />

+<br />

This is equivalent to what would be obta<strong>in</strong>ed from the variational pr<strong>in</strong>ciple<br />

+<br />

X<br />

=1<br />

<br />

<br />

<br />

=0 (5.41)<br />

X<br />

=0 (5.42)<br />

=1<br />

1 This textbook uses the symbol to designate a generalized coord<strong>in</strong>ate, and 0 to designate the correspond<strong>in</strong>g first derivative<br />

with respect to the <strong>in</strong>dependent variable, <strong>in</strong> order to differentiate the spatial coord<strong>in</strong>ates from the more powerful generalized<br />

coord<strong>in</strong>ates.


5.9. LAGRANGE MULTIPLIERS FOR HOLONOMIC CONSTRAINTS 123<br />

Equation 542 is equivalent to a variational problem for f<strong>in</strong>d<strong>in</strong>g the stationary value of 0<br />

Ã<br />

!<br />

X<br />

( 0 )= + =0 (5.43)<br />

where 0 is def<strong>in</strong>ed to be<br />

Ã<br />

!<br />

X<br />

0 ≡ + <br />

<br />

=1<br />

(5.44)<br />

The solution to equation 543 can be found us<strong>in</strong>g Euler’s differential equation 519 of variational calculus.<br />

At the extremum ( 0 )=0corresponds to follow<strong>in</strong>g contours of constant 0 which are <strong>in</strong> the surface that is<br />

perpendicular to the gradients of the terms <strong>in</strong> 0 . The Lagrange multiplier constants are required because,<br />

although these gradients are parallel at the extremum, the magnitudes of the gradients are not equal.<br />

The beauty of the Lagrange multipliers approach is that the auxiliary conditions do not have to be<br />

handled explicitly, s<strong>in</strong>ce they are handled automatically as additional free variables dur<strong>in</strong>g solution of<br />

Euler’s equations for a variational problem with + unknowns fit to + equations. That is, the <br />

variables are determ<strong>in</strong>ed by the variational procedure us<strong>in</strong>g the variational equations<br />

<br />

(0 ) − ( 0 )= <br />

<br />

0 <br />

(<br />

0 <br />

) − ( <br />

<br />

) −<br />

X<br />

<br />

<br />

<br />

<br />

=0 (5.45)<br />

simultaneously with the variables which are determ<strong>in</strong>ed by the variational equations<br />

Equation 545 usually is expressed as<br />

<br />

( 0 ) − ( 0 )=0 (5.46)<br />

<br />

0 <br />

( ) − <br />

( <br />

0 )+<br />

X<br />

<br />

<br />

<br />

<br />

=0 (5.47)<br />

The elegance of Lagrange multipliers is that a s<strong>in</strong>gle variational approach allows simultaneous determ<strong>in</strong>ation<br />

<br />

of all + unknowns. Chapter 62 shows that the forces of constra<strong>in</strong>t are given directly by the <br />

terms.<br />

5.7 Example: Two dependent variables coupled by one holonomic constra<strong>in</strong>t<br />

The powerful, and generally applicable, Lagrange multiplier technique is illustrated by consider<strong>in</strong>g the case<br />

of only two dependent variables, () and () with the function (() 0 ()()() 0 ; ) andwithone<br />

holonomic equation of constra<strong>in</strong>t coupl<strong>in</strong>g these two dependent variables. The extremum is given by requir<strong>in</strong>g<br />

∙µ <br />

− <br />

<br />

Z<br />

<br />

2<br />

µ<br />

= <br />

1<br />

0 + − <br />

<br />

0 <br />

with the constra<strong>in</strong>t expressed by the auxiliary condition<br />

¸<br />

=0<br />

()<br />

( ; ) =0<br />

()<br />

Note that the variations <br />

<br />

and<br />

<br />

are no longer <strong>in</strong>dependent because of the constra<strong>in</strong>t equation, thus the<br />

the two terms <strong>in</strong> the brackets of equation are not separately equal to zero at the extremum. However,<br />

differentiat<strong>in</strong>g the constra<strong>in</strong>t equation gives<br />

µ<br />

<br />

= <br />

+ <br />

<br />

=0 ()<br />

<br />

<br />

term applies because, for the <strong>in</strong>dependent variable,<br />

<br />

the auxiliary functions<br />

No <br />

<br />

=0 Introduce the neighbor<strong>in</strong>g paths by add<strong>in</strong>g<br />

( ) = ()+ 1 () ()<br />

( ) = ()+ 2 () ()


124 CHAPTER 5. CALCULUS OF VARIATIONS<br />

Insert the differentials of equations and , <strong>in</strong>to gives<br />

µ<br />

<br />

= 1()+ <br />

2()<br />

imply<strong>in</strong>g that<br />

Equation can be rewritten as<br />

Z 2<br />

∙µ <br />

1<br />

− <br />

<br />

0<br />

Z "<br />

2<br />

µ<br />

− <br />

<br />

0<br />

1<br />

<br />

<br />

2 () =− <br />

1 ()<br />

<br />

µ <br />

1 ()+<br />

µ <br />

−<br />

− <br />

=0 ( )<br />

− ¸<br />

<br />

0 2 () = 0<br />

<br />

#<br />

<br />

0 1 () = 0 ()<br />

Equation now conta<strong>in</strong>s only a s<strong>in</strong>gle arbitrary function 1 () that is not restricted by the constra<strong>in</strong>t. Thus<br />

the bracket <strong>in</strong> the <strong>in</strong>tegrand of equation must equal zero for the extremum. That is<br />

µ <br />

− <br />

<br />

µ −1<br />

<br />

0 =<br />

<br />

µ <br />

− <br />

<br />

<br />

<br />

µ −1<br />

<br />

0 ≡−()<br />

<br />

Now the left-hand side of this equation is only a function of and with respect to and 0 while the<br />

right-hand side is a function of and with respect to and 0 Becausebothsidesarefunctionsof then<br />

each side can be set equal to a function −() Thus the above equations can be written as<br />

<br />

0 − <br />

<br />

= ()<br />

<br />

<br />

<br />

0 − <br />

<br />

= ()<br />

<br />

<br />

()<br />

The complete solution of the three unknown functions. ()() and () is obta<strong>in</strong>ed by solv<strong>in</strong>g the two<br />

equations, , plus the equation of constra<strong>in</strong>t . The Lagrange multiplier () isrelatedtotheforceof<br />

constra<strong>in</strong>t. This example of two variables coupled by one holonomic constra<strong>in</strong>t conforms with the general<br />

relation for many variables and constra<strong>in</strong>ts given by equation 547.<br />

5.9.2 Integral equations of constra<strong>in</strong>t<br />

The constra<strong>in</strong>t equation also can be given <strong>in</strong> an <strong>in</strong>tegral form which is used frequently for isoperimetric<br />

problems. Consider a one dependent-variable isoperimetric problem, for f<strong>in</strong>d<strong>in</strong>g the curve = () such that<br />

the functional has an extremum, and the curve () satisfies boundary conditions such that ( 1 )= and<br />

( 2 )=. Thatis<br />

Z 2<br />

() = ( 0 ; ) (5.48)<br />

1<br />

is an extremum such that the fixed length of the perimeter satisfies the <strong>in</strong>tegral constra<strong>in</strong>t<br />

Z 2<br />

() = ( 0 ; ) = (5.49)<br />

1<br />

Analogous to (544) these two functionals can be comb<strong>in</strong>ed requir<strong>in</strong>g that<br />

( ) ≡ [ ()+()] = [ + ] =0 (5.50)<br />

1<br />

That is, it is an extremum for both () and the Lagrange multiplier . Thiseffectively <strong>in</strong>volves f<strong>in</strong>d<strong>in</strong>g the<br />

extremum path for the function ( ) = ( )+( ) where both () and are the m<strong>in</strong>imized<br />

variables. Therefore the curve () must satisfy the differential equation<br />

<br />

− ∙ <br />

+ − ¸<br />

=0 (5.51)<br />

<br />

0 <br />

0 <br />

Z 2


5.9. LAGRANGE MULTIPLIERS FOR HOLONOMIC CONSTRAINTS 125<br />

subject to the boundary conditions ( 1 )= ( 2 )= and () =.<br />

5.8 Example: Catenary<br />

One isoperimetric problem is the catenary which is the shape a uniform rope or cha<strong>in</strong> of fixed length <br />

that m<strong>in</strong>imizes the gravitational potential energy. Let the rope have a uniform mass per unit length of <br />

kg/m<br />

The gravitational potential energy is<br />

= <br />

Z 2<br />

1<br />

= <br />

Z 2<br />

1<br />

p 2 + 2 = <br />

The constra<strong>in</strong>t is that the length be a constant <br />

Z 2 Z 2 p<br />

= = 1+<br />

02<br />

<br />

1<br />

1<br />

Z 2<br />

1<br />

p 1+ 02 <br />

Thus the function is ( 0 ; ) = p 1+ 02 while the <strong>in</strong>tegral constra<strong>in</strong>t<br />

sets = p 1+ 02<br />

These need to be <strong>in</strong>serted <strong>in</strong>to the Euler equation (551) by def<strong>in</strong><strong>in</strong>g<br />

= + =( + ) p 1+ 02<br />

1 1<br />

The catenary<br />

Note that this case is one where <br />

<br />

=0and is a constant; also<br />

def<strong>in</strong><strong>in</strong>g = + then 0 = 0 Therefore the Euler’s equations can be written <strong>in</strong> the <strong>in</strong>tegral form<br />

− 0 <br />

0 = = constant<br />

Insert<strong>in</strong>g the relation = √ 1+ 02 gives<br />

p 1+ 02 − 0 0<br />

√ = 1+<br />

02<br />

where is an arbitrary constant. This simplifies to<br />

³ <br />

02 ´2<br />

= − 1<br />

<br />

The <strong>in</strong>tegral of this is<br />

µ + <br />

= cosh<br />

<br />

where and are arbitrary constants fixed by the locations of the two fixed ends of the rope.<br />

5.9 Example: The Queen Dido problem<br />

A famous constra<strong>in</strong>ed isoperimetric legend is that of Dido, first Queen of Carthage. Legend says that,<br />

when Dido landed <strong>in</strong> North Africa, she persuaded the local chief to sell her as much land as an oxhide could<br />

conta<strong>in</strong>. She cut an oxhide <strong>in</strong>to narrow strips and jo<strong>in</strong>ed them to make a cont<strong>in</strong>uous thread more than four<br />

kilometers <strong>in</strong> length which was sufficient to enclose the land adjo<strong>in</strong><strong>in</strong>g the coast on which Carthage was built.<br />

Her problem was to enclose the maximum area for a given perimeter. Let us assume that the coast l<strong>in</strong>e is<br />

straight and the ends of the thread are at ± on the coast l<strong>in</strong>e. The enclosed area is given by<br />

=<br />

Z +<br />

−<br />

<br />

The constra<strong>in</strong>t equation is that the total perimeter equals .<br />

Z <br />

−<br />

p<br />

1+<br />

02<br />

=


126 CHAPTER 5. CALCULUS OF VARIATIONS<br />

Thus we have that the functional ( 0 )= and ( 0 )= p 1+ 02 .<br />

and <br />

<br />

= √ 0<br />

Insert these <strong>in</strong>to the Euler-Lagrange equation (551) gives<br />

0 1+ 02<br />

" #<br />

1 − <br />

p 0<br />

1+<br />

02<br />

=0<br />

Then <br />

<br />

=1<br />

<br />

0<br />

=0 <br />

=0<br />

That is<br />

Integrate with respect to gives<br />

" #<br />

0<br />

p<br />

1+<br />

02<br />

= 1 <br />

0<br />

p<br />

1+<br />

02 = − <br />

where is a constant of <strong>in</strong>tegration. This can be rearranged to give<br />

The <strong>in</strong>tegral of this is<br />

Rearrang<strong>in</strong>g this gives<br />

0 ± ( − )<br />

= q<br />

2 − ( − ) 2<br />

q<br />

= ∓ 2 − ( − ) 2 + <br />

( − ) 2 +( − ) 2 = 2<br />

This is the equation of a circle centered at ( ). Sett<strong>in</strong>g the bounds to be (− 0) to ( 0) gives that<br />

= =0and the circle radius is Thus the length of the thread must be = . Assum<strong>in</strong>g that =4<br />

then =127 and Queen Dido could buy an area of 253 2 <br />

5.10 Geodesic<br />

Thegeodesicisdef<strong>in</strong>ed as the shortest path between two fixed po<strong>in</strong>ts for motion that is constra<strong>in</strong>ed to lie<br />

on a surface. <strong>Variational</strong> calculus provides a powerful approach for determ<strong>in</strong><strong>in</strong>g the equations of motion<br />

constra<strong>in</strong>ed to follow a geodesic.<br />

The use of variational calculus is illustrated by consider<strong>in</strong>g the geodesic constra<strong>in</strong>ed to follow the surface<br />

of a sphere of radius . As discussed <strong>in</strong> appendixq<br />

23, the element of path length on the surface of the<br />

sphere is given <strong>in</strong> spherical coord<strong>in</strong>ates as = 2 +(s<strong>in</strong>) 2 . Therefore the distance between two<br />

po<strong>in</strong>ts 1 and 2 is<br />

⎡s ⎤<br />

Z 2 µ 2<br />

= ⎣ <br />

+s<strong>in</strong> 2 ⎦ (5.52)<br />

<br />

1<br />

The function for ensur<strong>in</strong>g that be an extremum value uses<br />

p<br />

= 02 +s<strong>in</strong> 2 (5.53)<br />

where 0 = <br />

<br />

This is a case where<br />

<br />

lead<strong>in</strong>g to the result that<br />

This gives that<br />

This can be rewritten as<br />

=0and thus the <strong>in</strong>tegral form of Euler’s equation can be used<br />

p 02 +s<strong>in</strong> 2 − 0 <br />

0 p<br />

02 +s<strong>in</strong> 2 = constant = (5.54)<br />

p<br />

s<strong>in</strong> 2 = 02 +s<strong>in</strong> 2 (5.55)<br />

<br />

= 1 0 = csc 2 <br />

√<br />

1 − 2 csc 2 <br />

(5.56)


5.11. VARIATIONAL APPROACH TO CLASSICAL MECHANICS 127<br />

Solv<strong>in</strong>g for gives<br />

where<br />

That is<br />

Expand<strong>in</strong>g the s<strong>in</strong>e and cotangent gives<br />

µ cot <br />

=s<strong>in</strong> −1 + (5.57)<br />

<br />

≡ 1 − 2<br />

2 (5.58)<br />

cot = s<strong>in</strong> ( − ) (5.59)<br />

( cos ) s<strong>in</strong> s<strong>in</strong> − ( s<strong>in</strong> ) s<strong>in</strong> cos = cos (5.60)<br />

S<strong>in</strong>ce the brackets are constants, this can be written as<br />

( s<strong>in</strong> s<strong>in</strong> ) − ( s<strong>in</strong> cos ) =( cos ) (5.61)<br />

The terms <strong>in</strong> the brackets are just expressions for the rectangular coord<strong>in</strong>ates That is,<br />

− = (5.62)<br />

This is the equation of a plane pass<strong>in</strong>g through the center of the sphere. Thus the geodesic on a sphere<br />

is the path where a plane through the center <strong>in</strong>tersects the sphere as well as the <strong>in</strong>itial and f<strong>in</strong>al locations.<br />

This geodesic is called a great circle. Euler’s equation gives both the maximum and m<strong>in</strong>imum extremum<br />

path lengths for motion on this great circle.<br />

Chapter 17 discusses the geodesic <strong>in</strong> the four-dimensional space-time coord<strong>in</strong>ates that underlie the General<br />

Theory of Relativity. As a consequence, the use of the calculus of variations to determ<strong>in</strong>e the equations of<br />

motion for geodesics plays a pivotal role <strong>in</strong> the General Theory of Relativity.<br />

5.11 <strong>Variational</strong> approach to classical mechanics<br />

This chapter has <strong>in</strong>troduced the general pr<strong>in</strong>ciples of variational calculus needed for understand<strong>in</strong>g the Lagrangian<br />

and Hamiltonian approaches to classical mechanics. Although variational calculus was developed<br />

orig<strong>in</strong>ally for classical mechanics, now it has grown to be an important branch of mathematics with applications<br />

to many other fields outside of physics. The prologue of this book emphasized the dramatic differences<br />

between the differential vectorial approach of Newtonian mechanics, and the <strong>in</strong>tegral variational approaches<br />

of Lagrange and Hamiltonian mechanics. The Newtonian vectorial approach <strong>in</strong>volves solv<strong>in</strong>g Newton’s differential<br />

equations of motion that relate the force and momenta vectors. This requires knowledge of the<br />

time dependence of all the force vectors, <strong>in</strong>clud<strong>in</strong>g constra<strong>in</strong>t forces, act<strong>in</strong>g on the system which can be very<br />

complicated. Chapter 2 showed that the first-order time <strong>in</strong>tegrals, equations 210 216, relate the <strong>in</strong>itial and<br />

f<strong>in</strong>al total momenta without requir<strong>in</strong>g knowledge of the complicated <strong>in</strong>stantaneous forces act<strong>in</strong>g dur<strong>in</strong>g the<br />

collision of two bodies. Similarly, for conservative systems, the first-order spatial <strong>in</strong>tegral, equation 221,<br />

relates the <strong>in</strong>itial and f<strong>in</strong>al total energies to the net work done on the system without requir<strong>in</strong>g knowledge<br />

of the <strong>in</strong>stantaneous force vectors. The first-order spatial <strong>in</strong>tegral has the advantage that it is a scalar quantity,<br />

<strong>in</strong> contrast to time <strong>in</strong>tegrals which are vector quantities. These first-order <strong>in</strong>tegral relations are used<br />

frequently <strong>in</strong> Newtonian mechanics to derive solutions of the equations of motion that avoid hav<strong>in</strong>g to solve<br />

complicated differential equations of motion.<br />

This chapter has illustrated that variational pr<strong>in</strong>ciples provide a means of deriv<strong>in</strong>g more detailed <strong>in</strong>formation,<br />

such as the trajectories for the motion between given <strong>in</strong>itial and f<strong>in</strong>al conditions, by requir<strong>in</strong>g that<br />

scalar functionals have extrema values. For example, the solution of the brachistochrone problem determ<strong>in</strong>ed<br />

the trajectory hav<strong>in</strong>g the m<strong>in</strong>imum transit time, based on only the magnitudes of the k<strong>in</strong>etic and gravitational<br />

potential energies. Similarly, the catenary shape of a suspended cha<strong>in</strong> was derived by m<strong>in</strong>imiz<strong>in</strong>g the<br />

gravitational potential energy. The calculus of variations uses Euler’s equations to determ<strong>in</strong>e directly the<br />

differential equations of motion of the system that lead to the functional of <strong>in</strong>terest be<strong>in</strong>g stationary at an<br />

extremum. The Lagrangian and Hamiltonian variational approaches to classical mechanics are discussed<br />

<strong>in</strong> chapters 6 − 16. The broad range of applicability, the flexibility, and the power provided by variational<br />

approaches to classical mechanics and modern physics will be illustrated.


128 CHAPTER 5. CALCULUS OF VARIATIONS<br />

5.12 Summary<br />

Euler’s differential equation: The calculus of variations has been <strong>in</strong>troduced and Euler’s differential<br />

equation was derived. The calculus of variations reduces to vary<strong>in</strong>g the functions () where =1 2 3 ,<br />

such that the <strong>in</strong>tegral<br />

Z 2<br />

= [ ()(); 0 ] (516)<br />

1<br />

is an extremum, that is, it is a maximum or m<strong>in</strong>imum. Here is the <strong>in</strong>dependent variable, () are<br />

the dependent variables plus their first derivatives 0 ≡ <br />

The quantity [()0 (); ] has some given<br />

dependence on 0 and The calculus of variations <strong>in</strong>volves vary<strong>in</strong>g the functions () until a stationary<br />

value of is found which is presumed to be an extremum. It was shown that if the () are <strong>in</strong>dependent,<br />

then the extremum value of leads to <strong>in</strong>dependent Euler equations<br />

<br />

− <br />

<br />

0 =0 (519)<br />

where =1 2 3. This can be used to determ<strong>in</strong>e the functional form () that ensures that the <strong>in</strong>tegral<br />

= R 2<br />

1<br />

[() 0 (); ] is a stationary value, that is, presumably a maximum or m<strong>in</strong>imum value.<br />

Note that Euler’s equation <strong>in</strong>volves partial derivatives for the dependent variables 0 and the total<br />

derivative for the <strong>in</strong>dependent variable <br />

Euler’s <strong>in</strong>tegral equation: It was shown that if the function R 2<br />

1<br />

[ () 0 (); ] does not depend on<br />

the <strong>in</strong>dependent variable, then Euler’s differential equation can be written <strong>in</strong> an <strong>in</strong>tegral form. This <strong>in</strong>tegral<br />

form of Euler’s equation is especially useful when <br />

<br />

=0 that is, when does not depend explicitly on ,<br />

then the first <strong>in</strong>tegral of the Euler equation is a constant<br />

− 0 <br />

0 = constant<br />

(525)<br />

Constra<strong>in</strong>ed variational systems: Most applications <strong>in</strong>volve constra<strong>in</strong>ts on the motion. The equations<br />

of constra<strong>in</strong>t can be classified accord<strong>in</strong>g to whether the constra<strong>in</strong>ts are holonomic or non-holonomic, the time<br />

dependence of the constra<strong>in</strong>ts, and whether the constra<strong>in</strong>t forces are conservative.<br />

Generalized coord<strong>in</strong>ates <strong>in</strong> variational calculus: Independent generalized coord<strong>in</strong>ates can be chosen<br />

that are perpendicular to the rigid constra<strong>in</strong>t forces and therefore the constra<strong>in</strong>t does not contribute to the<br />

functional be<strong>in</strong>g m<strong>in</strong>imized. That is, the constra<strong>in</strong>ts are embedded <strong>in</strong>to the generalized coord<strong>in</strong>ates and thus<br />

the constra<strong>in</strong>ts can be ignored when deriv<strong>in</strong>g the variational solution.<br />

M<strong>in</strong>imal set of generalized coord<strong>in</strong>ates: If the constra<strong>in</strong>ts are holonomic then the holonomic<br />

equations of constra<strong>in</strong>t can be used to transform the coupled generalized coord<strong>in</strong>ates to = − <br />

<strong>in</strong>dependent generalized variables 0 . The generalized coord<strong>in</strong>ate method then uses Euler’s equations to<br />

determ<strong>in</strong>e these = − <strong>in</strong>dependent generalized coord<strong>in</strong>ates.<br />

<br />

− <br />

<br />

0 =0 (535)<br />

Lagrange multipliers for holonomic constra<strong>in</strong>ts: The Lagrange multipliers approach for variables,<br />

plus holonomic equations of constra<strong>in</strong>t, determ<strong>in</strong>es all + unknowns for the system. The holonomic<br />

forces of constra<strong>in</strong>t act<strong>in</strong>g on the variables, are related to the Lagrange multiplier terms () <br />

<br />

that<br />

are <strong>in</strong>troduced <strong>in</strong>to the Euler equations. That is,<br />

<br />

− <br />

<br />

<br />

0 <br />

where the holonomic equations of constra<strong>in</strong>t are given by<br />

+<br />

X<br />

<br />

() <br />

<br />

=0<br />

(548)<br />

( ; ) =0<br />

(538)<br />

The advantage of us<strong>in</strong>g the Lagrange multiplier approach is that the variational procedure simultaneously<br />

determ<strong>in</strong>es both the equations of motion for the variables plus the constra<strong>in</strong>t forces act<strong>in</strong>g on the<br />

system.


5.12. SUMMARY 129<br />

Problems<br />

1. F<strong>in</strong>d the extremal of the functional<br />

() =<br />

Z 2<br />

1<br />

˙ 2<br />

3 <br />

that satisfies (1) = 3 and (2) = 18. Show that this extremal provides the global m<strong>in</strong>imum of .<br />

2. Consider the use of equations of constra<strong>in</strong>t.<br />

(a) A particle is constra<strong>in</strong>ed to move on the surface of a sphere. What are the equations of constra<strong>in</strong>t for this<br />

system?<br />

(b) Adiskofmass and radius rolls without slipp<strong>in</strong>g on the outside surface of a half-cyl<strong>in</strong>der of radius<br />

5. What are the equations of constra<strong>in</strong>t for this system?<br />

(c) What are holonomic constra<strong>in</strong>ts? Which of the equations of constra<strong>in</strong>t that you found above are holonomic?<br />

(d) Equations of constra<strong>in</strong>t that do not explicitly conta<strong>in</strong> time are said to be scleronomic. Mov<strong>in</strong>g constra<strong>in</strong>ts<br />

are rheonomic. Are the equations of constra<strong>in</strong>t that you found above scleronomic or rheonomic?<br />

3. For each of the follow<strong>in</strong>g systems, describe the generalized coord<strong>in</strong>ates that would work best. There may be<br />

more than one answer for each system.<br />

(a) An <strong>in</strong>cl<strong>in</strong>ed plane of mass is slid<strong>in</strong>g on a smooth horizontal surface, while a particle of mass is<br />

slid<strong>in</strong>g on the smooth <strong>in</strong>cl<strong>in</strong>ed surface.<br />

(b) A disk rolls without slipp<strong>in</strong>g across a horizontal plane. The plane of the disk rema<strong>in</strong>s vertical, but it is<br />

free to rotate about a vertical axis.<br />

(c) A double pendulum consist<strong>in</strong>g of two simple pendula, with one pendulum suspended from the bob of the<br />

other. The two pendula have equal lengths and have bobs of equal mass. Both pendula are conf<strong>in</strong>ed to<br />

move<strong>in</strong>thesameplane.<br />

(d) Aparticleofmass is constra<strong>in</strong>ed to move on a circle of radius . The circle rotates <strong>in</strong> space about<br />

one po<strong>in</strong>t on the circle, which is fixed. The rotation takes place <strong>in</strong> the plane of the circle, with constant<br />

angular speed , <strong>in</strong> the absence of a gravitational force.<br />

(e) Aparticleofmass is attracted toward a given po<strong>in</strong>t by a force of magnitude 2 ,where is a constant.<br />

4. Look<strong>in</strong>g back at the systems <strong>in</strong> problem 3, which ones could have equations of constra<strong>in</strong>t? How would you<br />

classify the equations of constra<strong>in</strong>t (holonomic, scleronomic, rheonomic, etc.)?<br />

5. F<strong>in</strong>d the extremal of the functional<br />

() =<br />

Z <br />

0<br />

(2 s<strong>in</strong> − ˙ 2 )<br />

that satisfies () =() =0. Show that this extremal provides the global maximum of .<br />

q<br />

√<br />

6. F<strong>in</strong>d and describe the path = () for which the the <strong>in</strong>tegral 1+( 0 ) 2 is stationary.<br />

1<br />

7. F<strong>in</strong>d the dimensions of the parallelepiped of maximum volume circumscribed by a sphere of radius .<br />

8. Consider a s<strong>in</strong>gle loop of the cycloid hav<strong>in</strong>g a fixed value of as shown <strong>in</strong> the figure. A car released from<br />

rest at any po<strong>in</strong>t 0 anywhere on the track between and the lowest po<strong>in</strong>t ,thatis, 0 has a parameter<br />

0 0 <br />

Z 2


130 CHAPTER 5. CALCULUS OF VARIATIONS<br />

O<br />

x<br />

P<br />

0<br />

P<br />

y<br />

(a) Show that the time for the cart to slide from 0 to is given by the <strong>in</strong>tegral<br />

r Z r<br />

1 − cos <br />

( 0 → )=<br />

cos 0 − cos <br />

0<br />

(b) Prove that this time is equal to p which is <strong>in</strong>dependent of the position 0 <br />

(c) Expla<strong>in</strong> qualitatively how this surpris<strong>in</strong>g result can possibly be true.<br />

9. Consider a medium for which the refractive <strong>in</strong>dex = 2 where is a constant and is the distance from<br />

the orig<strong>in</strong>. Use Fermat’s Pr<strong>in</strong>ciple to f<strong>in</strong>d the path of a ray of light travell<strong>in</strong>g <strong>in</strong> a plane conta<strong>in</strong><strong>in</strong>g the orig<strong>in</strong>.<br />

H<strong>in</strong>t, use two-dimensional polar coord<strong>in</strong>ates with = () Show that the result<strong>in</strong>g path is a circle through<br />

the orig<strong>in</strong>.<br />

10. F<strong>in</strong>d the shortest path between the ( ) po<strong>in</strong>ts (0 −1 0) and (0 1 0) on the conical surface<br />

=1− p 2 + 2<br />

What is the length of this path? Note that this is the shortest mounta<strong>in</strong> path around a volcano.<br />

11. Show that the geodesic on the surface of a right circular cyl<strong>in</strong>der is a segment of a helix.


Chapter 6<br />

Lagrangian dynamics<br />

6.1 Introduction<br />

Newtonian mechanics is based on vector observables such as momentum and force, and Newton’s equations<br />

of motion can be derived if the forces are known. Newtonian mechanics becomes difficult to apply for manybody<br />

systems that <strong>in</strong>volve constra<strong>in</strong>t forces. The alternative algebraic Lagrangian mechanics approach is<br />

based on the concept of scalar energies which circumvent many of the difficulties <strong>in</strong> handl<strong>in</strong>g constra<strong>in</strong>t forces<br />

and many-body systems.<br />

The Lagrangian approach to classical dynamics is based on the calculus of variations <strong>in</strong>troduced <strong>in</strong> chapter<br />

5. It was shown that the calculus of variations determ<strong>in</strong>es the function () such that the scalar functional<br />

=<br />

Z 2<br />

1<br />

X<br />

<br />

[ () 0 (); ] (6.1)<br />

is an extremum, that is, a maximum or m<strong>in</strong>imum. Here is the <strong>in</strong>dependent variable, () are the <br />

dependent variables, and their derivatives 0 ≡ <br />

where =1 2 3 The function [ () 0 (); ] has<br />

an assumed dependence on 0 and The calculus of variations determ<strong>in</strong>es the functional dependence<br />

of the dependent variables () on the <strong>in</strong>dependent variable that is needed to ensure that is an<br />

extremum. For <strong>in</strong>dependent variables, has a stationary po<strong>in</strong>t, which is presumed to be an extremum,<br />

that is determ<strong>in</strong>ed by solution of Euler’s differential equations<br />

<br />

<br />

0 − =0 (6.2)<br />

<br />

If the coord<strong>in</strong>ates () are <strong>in</strong>dependent, then the Euler equations, (62), for each coord<strong>in</strong>ate are <strong>in</strong>dependent.<br />

However, for constra<strong>in</strong>ed motion, the constra<strong>in</strong>ts lead to auxiliary conditions that correlate the<br />

coord<strong>in</strong>ates. As shown <strong>in</strong> chapter 5 a transformation to <strong>in</strong>dependent generalized coord<strong>in</strong>ates can be made<br />

such that the correlations <strong>in</strong>duced by the constra<strong>in</strong>t forces are embedded <strong>in</strong>to the choice of the <strong>in</strong>dependent<br />

generalized coord<strong>in</strong>ates. The use of generalized coord<strong>in</strong>ates <strong>in</strong> Lagrangian mechanics simplifies derivation of<br />

the equations of motion for constra<strong>in</strong>ed systems. For example, for a system of coord<strong>in</strong>ates, that <strong>in</strong>volves<br />

holonomic constra<strong>in</strong>ts, there are = − <strong>in</strong>dependent generalized coord<strong>in</strong>ates. For such holonomic<br />

constra<strong>in</strong>ed motion, it will be shown that the Euler equations can be solved us<strong>in</strong>g either of the follow<strong>in</strong>g<br />

three alternative ways.<br />

1) The m<strong>in</strong>imal set of generalized coord<strong>in</strong>ates approach <strong>in</strong>volves f<strong>in</strong>d<strong>in</strong>g a set of = − <strong>in</strong>dependent<br />

generalized coord<strong>in</strong>ates that satisfy the assumptions underly<strong>in</strong>g (62). These generalized coord<strong>in</strong>ates<br />

can be determ<strong>in</strong>ed if the equations of constra<strong>in</strong>t are holonomic, that is, related by algebraic equations of<br />

constra<strong>in</strong>t<br />

( ; ) =0 (6.3)<br />

where =1 2 3 These equations uniquely determ<strong>in</strong>e the relationship between the correlated coord<strong>in</strong>ates.<br />

This method has the advantage that it reduces the system of coord<strong>in</strong>ates, subject to constra<strong>in</strong>ts,<br />

to = − <strong>in</strong>dependent generalized coord<strong>in</strong>ates which reduces the dimension of the problem to be solved.<br />

However, it does not explicitly determ<strong>in</strong>e the forces of constra<strong>in</strong>t which are effectively swept under the rug.<br />

131


132 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

2) The Lagrange multipliers approach takes account of the correlation between the coord<strong>in</strong>ates and<br />

holonomic constra<strong>in</strong>ts by <strong>in</strong>troduc<strong>in</strong>g the Lagrange multipliers (). These generalized coord<strong>in</strong>ates <br />

are correlated by the holonomic constra<strong>in</strong>ts.<br />

<br />

<br />

0 <br />

− <br />

<br />

=<br />

X<br />

<br />

() <br />

<br />

(6.4)<br />

where =1 2 3. The Lagrange multiplier approach has the advantage that Euler’s calculus of variations<br />

automatically use the Lagrange equations, plus the equations of constra<strong>in</strong>t, to explicitly determ<strong>in</strong>e both<br />

the coord<strong>in</strong>ates and the forces of constra<strong>in</strong>t which are related to the Lagrange multipliers as given<br />

<strong>in</strong> equation (64). Chapter 62 shows that the P <br />

() <br />

<br />

terms are directly related to the holonomic<br />

forces of constra<strong>in</strong>t.<br />

3) The generalized force approach <strong>in</strong>corporates the forces of constra<strong>in</strong>t explicitly as will be shown <strong>in</strong><br />

chapter 654. Incorporat<strong>in</strong>g the constra<strong>in</strong>t forces explicitly allows use of holonomic, non-holonomic, and<br />

non-conservative constra<strong>in</strong>t forces.<br />

Understand<strong>in</strong>g the Lagrange formulation of classical mechanics is facilitated by use of a simple nonrigorous<br />

plausibility approach that is based on Newton’s laws of motion. This <strong>in</strong>troductory plausibility approach<br />

will be followed by two more rigorous derivations of the Lagrangian formulation developed us<strong>in</strong>g either<br />

d’Alembert Pr<strong>in</strong>ciple or Hamiltons Pr<strong>in</strong>ciple. These better elucidate the physics underly<strong>in</strong>g the Lagrange<br />

and Hamiltonian analytic representations of classical mechanics. In 1788 Lagrange derived his equations of<br />

motion us<strong>in</strong>g the differential d’Alembert Pr<strong>in</strong>ciple, that extends to dynamical systems the Bernoulli Pr<strong>in</strong>ciple<br />

of <strong>in</strong>f<strong>in</strong>itessimal virtual displacements and virtual work. The other approach, developed <strong>in</strong> 1834, usesthe<br />

<strong>in</strong>tegral Hamilton’s Pr<strong>in</strong>ciple to derive the Lagrange equations. Hamilton’s Pr<strong>in</strong>ciple is discussed <strong>in</strong> more<br />

detail <strong>in</strong> chapter 9 Euler’s variational calculus underlies d’Alembert’s Pr<strong>in</strong>ciple and Hamilton’s Pr<strong>in</strong>ciple<br />

s<strong>in</strong>ce both are based on the philosophical belief that the laws of nature prefer economy of motion. Chapters<br />

62 − 65 show that both d’Alembert’s Pr<strong>in</strong>ciple and Hamilton’s Pr<strong>in</strong>ciple lead to the Euler-Lagrange<br />

equations. This will be followed by a series of examples that illustrate the use of Lagrangian mechanics <strong>in</strong><br />

classical mechanics.<br />

6.2 Newtonian plausibility argument for Lagrangian mechanics<br />

Insight <strong>in</strong>to the physics underly<strong>in</strong>g Lagrange mechanics is given by show<strong>in</strong>g the direct relationship between<br />

Newtonian and Lagrangian mechanics. The variational approaches to classical mechanics exploit the firstorder<br />

spatial <strong>in</strong>tegral of the force, equation 217 which equals the work done between the <strong>in</strong>itial and f<strong>in</strong>al<br />

conditions. The work done is a simple scalar quantity that depends on the <strong>in</strong>itial and f<strong>in</strong>al location for<br />

conservative forces. Newton’s equation of motion is<br />

The k<strong>in</strong>etic energy is given by<br />

F = p<br />

<br />

(6.5)<br />

= 1 2 2 = p · p<br />

2 = 2 <br />

2 + 2 <br />

2 + 2 <br />

2<br />

It can be seen that<br />

<br />

˙ = (6.6)<br />

and<br />

<br />

˙ = <br />

= (6.7)<br />

Consider that the force, act<strong>in</strong>g on a mass is arbitrarily separated <strong>in</strong>to two components, one part that<br />

is conservative, and thus can be written as the gradient of a scalar potential , plus the excluded part of<br />

the force, . The excluded part of the force could <strong>in</strong>clude non-conservative frictional forces as well<br />

as forces of constra<strong>in</strong>t which may be conservative or non-conservative. This separation allows the force to<br />

be written as<br />

F = −∇ + F (6.8)


6.2. NEWTONIAN PLAUSIBILITY ARGUMENT FOR LAGRANGIAN MECHANICS 133<br />

Along each of the axes,<br />

<br />

= − + <br />

˙ <br />

(6.9)<br />

<br />

Equation (69) can be extended by transform<strong>in</strong>g the cartesian coord<strong>in</strong>ate to the generalized coord<strong>in</strong>ates<br />

<br />

Def<strong>in</strong>e the standard Lagrangian to be the difference between the k<strong>in</strong>etic energy and the potential energy,<br />

which can be written <strong>in</strong> terms of the generalized coord<strong>in</strong>ates as<br />

( ˙ ) ≡ ( ˙ ) − ( ) (6.10)<br />

Assume that the potential is only a function of the generalized coord<strong>in</strong>ates that is <br />

˙ <br />

=0 then<br />

<br />

= + = <br />

(6.11)<br />

˙ ˙ ˙ ˙ <br />

Us<strong>in</strong>g the above equations allows Newton’s equation of motion (69) to be expressed as<br />

The excluded force <br />

<br />

excluded forces as given by<br />

<br />

− = <br />

˙ <br />

(6.12)<br />

<br />

can be partitioned <strong>in</strong>to a holonomic constra<strong>in</strong>t force <br />

<br />

plus any rema<strong>in</strong><strong>in</strong>g<br />

<br />

<br />

= <br />

<br />

+ (6.13)<br />

A comparison of equations (612 613) and (64) shows that the holonomic constra<strong>in</strong>t forces <br />

<br />

that are<br />

conta<strong>in</strong>ed <strong>in</strong> the excluded force can be identified with the Lagrange multiplier term <strong>in</strong> equation 64.<br />

<br />

<br />

≡<br />

X<br />

<br />

() <br />

<br />

(6.14)<br />

That is the Lagrange multiplier terms can be used to account for holonomic constra<strong>in</strong>t forces <br />

<br />

. Thus<br />

equation 612 can be written as<br />

<br />

− X<br />

= () <br />

+ <br />

˙ <br />

(6.15)<br />

<br />

<br />

where the Lagrange multiplier term accounts for holonomic constra<strong>in</strong>t forces, and <br />

<br />

<strong>in</strong>cludes all the<br />

rema<strong>in</strong><strong>in</strong>g forces that are not accounted for by the scalar potential , or the Lagrange multiplier terms <br />

<br />

.<br />

For holonomic, conservative forces it is possible to absorb all the forces <strong>in</strong>to the potential plus the<br />

Lagrange multiplier term, that is <br />

<br />

=0 Moreover, the use of a m<strong>in</strong>imal set of generalized coord<strong>in</strong>ates<br />

allows the holonomic constra<strong>in</strong>t forces to be ignored by explicitly reduc<strong>in</strong>g the number of coord<strong>in</strong>ates from<br />

dependent coord<strong>in</strong>ates to = − <strong>in</strong>dependent generalized coord<strong>in</strong>ates. That is, the correlations due<br />

to the constra<strong>in</strong>t forces are embedded <strong>in</strong>to the generalized coord<strong>in</strong>ates. Then equation 615 reduces to the<br />

basic Euler differential equations.<br />

<br />

− =0 (6.16)<br />

˙ <br />

Note that equation 616 is identical to Euler’s equation 534, if the <strong>in</strong>dependent variable is replaced<br />

by time . Thus Newton’s equation of motion are equivalent to m<strong>in</strong>imiz<strong>in</strong>g the action <strong>in</strong>tegral = R 2<br />

1<br />

,<br />

that is<br />

Z 2<br />

= ( ˙ ; ) =0 (6.17)<br />

1<br />

which is Hamilton’s Pr<strong>in</strong>ciple. Hamilton’s Pr<strong>in</strong>ciple underlies many aspects of physics and as discussed <strong>in</strong><br />

chapter 9, and is used as the start<strong>in</strong>g po<strong>in</strong>t for develop<strong>in</strong>g classical mechanics. Hamilton’s Pr<strong>in</strong>ciple was<br />

postulated 46 years after Lagrange <strong>in</strong>troduced Lagrangian mechanics.<br />

The above plausibility argument, which is based on Newtonian mechanics, illustrates the close connection<br />

between the vectorial Newtonian mechanics and the algebraic Lagrangian mechanics approaches to classical<br />

mechanics.


134 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

6.3 Lagrange equations from d’Alembert’s Pr<strong>in</strong>ciple<br />

6.3.1 d’Alembert’s Pr<strong>in</strong>ciple of Virtual Work<br />

The Pr<strong>in</strong>ciple of Virtual Work provides a basis for a rigorous derivation of Lagrangian mechanics. Bernoulli<br />

<strong>in</strong>troduced the concept of virtual <strong>in</strong>f<strong>in</strong>itessimal displacement of a system mentioned <strong>in</strong> chapter 591. This<br />

refers to a change <strong>in</strong> the configuration of the system as a result of any arbitrary <strong>in</strong>f<strong>in</strong>itessimal <strong>in</strong>stantaneous<br />

change of the coord<strong>in</strong>ates r that is consistent with the forces and constra<strong>in</strong>ts imposed on the system at<br />

the <strong>in</strong>stant . Lagrange’s symbol is used to designate a virtual displacement which is called “virtual” to<br />

imply that there is no change <strong>in</strong> time , i.e. =0. This dist<strong>in</strong>guishes it from an actual displacement r of<br />

body dur<strong>in</strong>g a time <strong>in</strong>terval when the forces and constra<strong>in</strong>ts may change.<br />

Suppose that the system of particles is <strong>in</strong> equilibrium, that is, the total force on each particle is<br />

zero. The virtual work done by the force F mov<strong>in</strong>gadistancer is given by the dot product F · r .For<br />

equilibrium, the sum of all these products for the bodies also must be zero<br />

X<br />

F · r =0 (6.18)<br />

Decompos<strong>in</strong>g the force F on particle <strong>in</strong>to applied forces F and constra<strong>in</strong>t forces f gives<br />

<br />

<br />

X<br />

X<br />

F · r + f · r =0 (6.19)<br />

The second term <strong>in</strong> equation 619 can be ignored if the virtual work due to the constra<strong>in</strong>t forces is zero.<br />

This is rigorously true for rigid bodies and is valid for any forces of constra<strong>in</strong>t where the constra<strong>in</strong>t forces<br />

are perpendicular to the constra<strong>in</strong>t surface and the virtual displacement is tangent to this surface. Thus if<br />

the constra<strong>in</strong>t forces do no work, then (619) reduces to<br />

<br />

X<br />

F · r =0 (6.20)<br />

<br />

This relation is the Bernoulli’s Pr<strong>in</strong>ciple of Static Virtual Work and is used to solve problems <strong>in</strong> statics.<br />

Bernoulli <strong>in</strong>troduced dynamics by us<strong>in</strong>g Newton’s Law to related force and momentum.<br />

F = ṗ (6.21)<br />

Equation (621) can be rewritten as<br />

F − ṗ =0 (6.22)<br />

In 1742, d’Alembert developed the Pr<strong>in</strong>ciple of Dynamic Virtual Work <strong>in</strong> the form<br />

Us<strong>in</strong>g equations (619) plus (623) gives<br />

<br />

X<br />

(F − ṗ ) · r =0 (6.23)<br />

<br />

X<br />

X<br />

(F − ṗ ) · r + f · r =0 (6.24)<br />

For the special case where the forces of constra<strong>in</strong>t are zero, then equation 624 reduces to d’Alembert’s<br />

Pr<strong>in</strong>ciple<br />

X<br />

(F − ṗ ) · r =0 (6.25)<br />

<br />

d’Alembert’s Pr<strong>in</strong>ciple, by a stroke of genius, cleverly transforms the pr<strong>in</strong>ciple of virtual work from the realm<br />

of statics to dynamics. Application of virtual work to statics primarily leads to algebraic equations between<br />

the forces, whereas d’Alembert’s pr<strong>in</strong>ciple applied to dynamics leads to differential equations.


6.3. LAGRANGE EQUATIONS FROM D’ALEMBERT’S PRINCIPLE 135<br />

6.3.2 Transformation to generalized coord<strong>in</strong>ates<br />

In classical mechanical systems the coord<strong>in</strong>ates r usually are not <strong>in</strong>dependent due to the forces of constra<strong>in</strong>t<br />

and the constra<strong>in</strong>t-force energy contributes to equation 624. These problems can be elim<strong>in</strong>ated by express<strong>in</strong>g<br />

d’Alembert’s Pr<strong>in</strong>ciple <strong>in</strong> terms of virtual displacements of <strong>in</strong>dependent generalized coord<strong>in</strong>ates of the<br />

system for which the constra<strong>in</strong>t force term P <br />

f <br />

· q =0. Then the <strong>in</strong>dividual variational coefficients <br />

are <strong>in</strong>dependent and (F − ṗ ) · q =0can be equated to zero for each value of .<br />

The transformation of the -body system to <strong>in</strong>dependent generalized coord<strong>in</strong>ates can be expressed<br />

as<br />

r = r ( 1 2 3 ) (6.26)<br />

Assum<strong>in</strong>g <strong>in</strong>dependent coord<strong>in</strong>ates, then the velocity v can be written <strong>in</strong> terms of general coord<strong>in</strong>ates <br />

us<strong>in</strong>g the cha<strong>in</strong> rule for partial differentiation.<br />

v ≡ r <br />

= X r <br />

˙ + r <br />

<br />

<br />

(6.27)<br />

The arbitrary virtual displacement r can be related to the virtual displacement of the generalized coord<strong>in</strong>ate<br />

by<br />

X r <br />

r = (6.28)<br />

<br />

<br />

Note that by def<strong>in</strong>ition, a virtual displacement considers only displacements of the coord<strong>in</strong>ates, and no time<br />

variation is <strong>in</strong>volved.<br />

The above transformations can be used to express d’Alembert’s dynamical pr<strong>in</strong>ciple of virtual work <strong>in</strong><br />

generalized coord<strong>in</strong>ates. Thus the first term <strong>in</strong> d’Alembert’s Dynamical Pr<strong>in</strong>ciple, (625) becomes<br />

X<br />

F · r =<br />

<br />

X<br />

<br />

F <br />

· r <br />

<br />

=<br />

X<br />

(6.29)<br />

<br />

where are called components of the generalized force, 1 def<strong>in</strong>ed as<br />

X<br />

≡ F · r <br />

(6.30)<br />

<br />

<br />

Note that just as the generalized coord<strong>in</strong>ates need not have the dimensions of length, so the do not<br />

necessarily have the dimensions of force, but the product must have the dimensions of work. For<br />

example, couldbetorqueand could be the correspond<strong>in</strong>g <strong>in</strong>f<strong>in</strong>itessimal rotation angle.<br />

The second term <strong>in</strong> d’Alembert’s Pr<strong>in</strong>ciple (625) can be transformed us<strong>in</strong>g equation 628<br />

Ã<br />

X<br />

X<br />

<br />

!<br />

X<br />

ṗ · r = ¨r · r = ¨r · r <br />

(6.31)<br />

<br />

<br />

<br />

<br />

The right-hand side of (631) can be rewritten as<br />

Ã<br />

X <br />

!<br />

¨r · r X<br />

½ µ<br />

<br />

<br />

= ṙ · r <br />

<br />

− ṙ · µ ¾ r<br />

(6.32)<br />

<br />

<br />

Note that equation (627) gives that<br />

v <br />

= r <br />

(6.33)<br />

˙ <br />

therefore the first right-hand term <strong>in</strong> (632) can be written as<br />

<br />

<br />

µ<br />

ṙ · r <br />

<br />

<br />

= <br />

<br />

µ<br />

v · v <br />

<br />

˙ <br />

1 This proof, plus the notation, conform with that used by Goldste<strong>in</strong> [Go50] and by other texts on classical mechanics.<br />

(6.34)


136 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

The second right-hand term <strong>in</strong> (632) can be rewritten by <strong>in</strong>terchang<strong>in</strong>g the order of the differentiation with<br />

respect to and <br />

µ <br />

r<br />

= v <br />

(6.35)<br />

<br />

Substitut<strong>in</strong>g (634) and (635) <strong>in</strong>to (632) gives<br />

Ã<br />

X<br />

<br />

!<br />

X<br />

ṗ · r = ¨r · r X<br />

½ µ<br />

<br />

<br />

= v · v <br />

<br />

− v · v ¾<br />

<br />

(6.36)<br />

˙ <br />

<br />

<br />

Insert<strong>in</strong>g (629) and (636) <strong>in</strong>to d’Alembert’s Pr<strong>in</strong>ciple (625) leads to the relation<br />

( Ã Ã<br />

X<br />

X<br />

(F X<br />

− ṗ ) · r = −<br />

˙<br />

<br />

<br />

<br />

<br />

<br />

!!<br />

1<br />

2 <br />

2 −<br />

à X <br />

<br />

<br />

! )<br />

1<br />

2 <br />

2 − =0 (6.37)<br />

The P <br />

<br />

1 2 2 term can be identified with the system k<strong>in</strong>etic energy . Thus d’Alembert Pr<strong>in</strong>ciple reduces<br />

to the relation<br />

X<br />

∙½ <br />

<br />

<br />

µ <br />

˙ <br />

<br />

− <br />

<br />

¾<br />

− ¸<br />

=0 (6.38)<br />

For cartesian coord<strong>in</strong>ates is a function only of velocities ( ˙ ˙ ˙) and thus the term <br />

<br />

as discussed <strong>in</strong> appendix 22, for curvil<strong>in</strong>ear coord<strong>in</strong>ates <br />

<br />

=0 However,<br />

6=0due to the curvature of the coord<strong>in</strong>ates<br />

as is illustrated for polar coord<strong>in</strong>ates where v = ˙ˆr + ˙ˆθ.<br />

If all the generalized coord<strong>in</strong>ates are <strong>in</strong>dependent, thenequation638 implies that the term <strong>in</strong> the<br />

square brackets is zero for each <strong>in</strong>dividual value of . This leads to the basicEuler-Lagrangeequationsof<br />

motion for each of the <strong>in</strong>dependent generalized coord<strong>in</strong>ates<br />

½ <br />

<br />

µ <br />

˙ <br />

<br />

− <br />

<br />

¾<br />

= (6.39)<br />

where ≥ ≥ 1. That is, this leads to Euler-Lagrange equations of motion for the generalized forces .<br />

As discussed <strong>in</strong> chapter 58 when holonomic constra<strong>in</strong>t forces apply, it is possible to reduce the system<br />

to = − <strong>in</strong>dependent generalized coord<strong>in</strong>ates for which equation 625 applies.<br />

In 1687 Leibniz proposed m<strong>in</strong>imiz<strong>in</strong>g the time <strong>in</strong>tegral of his “vis viva”, which equals 2 That is,<br />

Z 2<br />

=0 (6.40)<br />

1<br />

The variational equation 639 accomplishes the m<strong>in</strong>imization of equation 640. It is remarkable that Leibniz<br />

anticipated the basic variational concept prior to the birth of the developers of Lagrangian mechanics, i.e.,<br />

d’Alembert, Euler, Lagrange, and Hamilton.<br />

6.3.3 Lagrangian<br />

The handl<strong>in</strong>g of both conservative and non-conservative generalized forces is best achieved by assum<strong>in</strong>g<br />

that the generalized force = P <br />

F · ¯r<br />

<br />

can be partitioned <strong>in</strong>to a conservative velocity-<strong>in</strong>dependent term,<br />

that can be expressed <strong>in</strong> terms of the gradient of a scalar potential, −∇ plus an excluded generalized force<br />

<br />

which conta<strong>in</strong>s the non-conservative, velocity-dependent, and all the constra<strong>in</strong>t forces not explicitly<br />

<strong>in</strong>cluded <strong>in</strong> the potential .Thatis,<br />

= −∇ + <br />

(6.41)<br />

Insert<strong>in</strong>g (641) <strong>in</strong>to (638) and assum<strong>in</strong>g that the potential is velocity <strong>in</strong>dependent, allows(638) to be<br />

rewritten as<br />

X<br />

∙½ µ <br />

¾ ¸<br />

( − ) ( − )<br />

− − <br />

=0 (6.42)<br />

˙


6.4. LAGRANGE EQUATIONS FROM HAMILTON’S ACTION PRINCIPLE 137<br />

The def<strong>in</strong>ition of the Standard Lagrangian is<br />

then (642) can be written as<br />

X<br />

∙½ <br />

<br />

<br />

≡ − (6.43)<br />

µ <br />

− ¾ ¸<br />

− <br />

=0 (6.44)<br />

˙ <br />

Note that equation (644) conta<strong>in</strong>s the basic Euler-Lagrange equation (638) as a special case when =0.<br />

In addition, note that if all the generalized coord<strong>in</strong>ates are <strong>in</strong>dependent, then the square bracket terms are<br />

zero for each value of which leads to the general Euler-Lagrange equations of motion<br />

½ µ <br />

− ¾<br />

= <br />

(6.45)<br />

˙ <br />

where ≥ ≥ 1.<br />

Chapter 653 will show that the holonomic constra<strong>in</strong>t forces can be factored out of the generalized force<br />

term <br />

which simplifies derivation of the equations of motion us<strong>in</strong>g Lagrangian mechanics. The general<br />

Euler-Lagrange equations of motion are used extensively <strong>in</strong> classical mechanics because conservative forces<br />

play a ubiquitous role <strong>in</strong> classical mechanics.<br />

6.4 Lagrange equations from Hamilton’s Action Pr<strong>in</strong>ciple<br />

Hamilton published two papers <strong>in</strong> 1834 and 1835 announc<strong>in</strong>g a fundamental new dynamical pr<strong>in</strong>ciple that<br />

underlies both Lagrangian and Hamiltonian mechanics. Hamilton was seek<strong>in</strong>g a theory of optics when he<br />

developed Hamilton’s Action Pr<strong>in</strong>ciple, plus the field of Hamiltonian mechanics, both of which play a crucial<br />

role <strong>in</strong> classical mechanics and modern physics. Hamilton’s Action Pr<strong>in</strong>ciple states “dynamicalsystems<br />

follow paths that m<strong>in</strong>imize the time <strong>in</strong>tegral of the Lagrangian”. That is, the action functional <br />

=<br />

Z 2<br />

1<br />

(q ˙q) (6.46)<br />

has a m<strong>in</strong>imum value for the correct path of motion. Hamilton’s Action Pr<strong>in</strong>ciple can be written <strong>in</strong><br />

terms of a virtual <strong>in</strong>f<strong>in</strong>itessimal displacement as<br />

Z 2<br />

= =0 (6.47)<br />

1<br />

<strong>Variational</strong> calculus therefore implies that a system of <strong>in</strong>dependent generalized coord<strong>in</strong>ates must satisfy<br />

the basic Lagrange-Euler equations<br />

<br />

− =0 (6.48)<br />

˙ <br />

Note that for <br />

=0 this is the same as equation 645 which was derived us<strong>in</strong>g d’Alembert’s Pr<strong>in</strong>ciple.<br />

This discussion has shown that Euler’s variational differential equation underlies both the differential variational<br />

d’Alembert Pr<strong>in</strong>ciple, and the more fundamental <strong>in</strong>tegral Hamilton’s Action Pr<strong>in</strong>ciple. As discussed<br />

<strong>in</strong> chapter 92, Hamilton’s Pr<strong>in</strong>ciple of Stationary Action adds a fundamental new dimension to classical<br />

mechanics which leads to derivation of both Lagrangian and Hamiltonian mechanics. That is, both Hamilton’s<br />

Action Pr<strong>in</strong>ciple, and d’Alembert’s Pr<strong>in</strong>ciple, can be used to derive Lagrangian mechanics lead<strong>in</strong>g to<br />

the most general Lagrange equations that are applicable to both holonomic and non-holonomic constra<strong>in</strong>ts,<br />

as well as conservative and non-conservative systems. In addition, Chapter 62 presented a plausibility argument<br />

show<strong>in</strong>g that Lagrangian mechanics can be justified based on Newtonian mechanics. Hamilton’s<br />

Action Pr<strong>in</strong>ciple, and d’Alembert’s Pr<strong>in</strong>ciple, can be expressed<strong>in</strong>termsofgeneralizedcoord<strong>in</strong>ateswhichis<br />

much broader <strong>in</strong> scope than the equations of motion implied us<strong>in</strong>g Newtonian mechanics.


138 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

6.5 Constra<strong>in</strong>ed systems<br />

The motion for systems subject to constra<strong>in</strong>ts is difficult to calculate us<strong>in</strong>g Newtonian mechanics because<br />

all the unknown constra<strong>in</strong>t forces must be <strong>in</strong>cluded explicitly with the active forces <strong>in</strong> order to determ<strong>in</strong>e<br />

the equations of motion. Lagrangian mechanics avoids these difficulties by allow<strong>in</strong>g selection of <strong>in</strong>dependent<br />

generalized coord<strong>in</strong>ates that <strong>in</strong>corporate the correlated motion <strong>in</strong>duced by the constra<strong>in</strong>t forces. This allows<br />

the constra<strong>in</strong>t forces act<strong>in</strong>g on the system to be ignored by reduc<strong>in</strong>g the system to a m<strong>in</strong>imal set of generalized<br />

coord<strong>in</strong>ates. The holonomic constra<strong>in</strong>t forces can be determ<strong>in</strong>ed us<strong>in</strong>g the Lagrange multiplier approach, or<br />

all constra<strong>in</strong>t forces can be determ<strong>in</strong>ed by <strong>in</strong>clud<strong>in</strong>g them as generalized forces, as described below.<br />

6.5.1 Choice of generalized coord<strong>in</strong>ates<br />

As discussed <strong>in</strong> chapter 58, theflexibility and freedom for selection of generalized coord<strong>in</strong>ates is a considerable<br />

advantage of Lagrangian mechanics when handl<strong>in</strong>g constra<strong>in</strong>ed systems. The generalized coord<strong>in</strong>ates<br />

can be any set of <strong>in</strong>dependent variables that completely specify the scalar action functional, equation 646.<br />

The generalized coord<strong>in</strong>ates are not required to be orthogonal as is required when us<strong>in</strong>g the vectorial Newtonian<br />

approach. The secret to us<strong>in</strong>g generalized coord<strong>in</strong>ates is to select coord<strong>in</strong>ates that are perpendicular<br />

to the constra<strong>in</strong>t forces so that the constra<strong>in</strong>t forces do no work. Moreover, if the constra<strong>in</strong>ts are rigid, then<br />

the constra<strong>in</strong>t forces do no work <strong>in</strong> the direction of the constra<strong>in</strong>t force. As a consequence, the constra<strong>in</strong>t<br />

forces do not contribute to the action <strong>in</strong>tegral and thus the P <br />

f <br />

· r term <strong>in</strong> equation 619 can be omitted<br />

from the action <strong>in</strong>tegral. Generalized coord<strong>in</strong>ates allow reduc<strong>in</strong>g the number of unknowns from to<br />

= − when the system has holonomic constra<strong>in</strong>ts. In addition, generalized coord<strong>in</strong>ates facilitate<br />

us<strong>in</strong>g both the Lagrange multipliers, and the generalized forces, approaches for determ<strong>in</strong><strong>in</strong>g the constra<strong>in</strong>t<br />

forces.<br />

6.5.2 M<strong>in</strong>imal set of generalized coord<strong>in</strong>ates<br />

The set of generalized coord<strong>in</strong>ates are used to describe the motion of the system. No restrictions have<br />

been placed on the nature of the constra<strong>in</strong>ts other than they are workless for a virtual displacement. If the<br />

constra<strong>in</strong>ts are holonomic, then it is possible to f<strong>in</strong>d sets of = − <strong>in</strong>dependent generalized coord<strong>in</strong>ates<br />

that conta<strong>in</strong> the constra<strong>in</strong>t conditions implicitly <strong>in</strong> the transformation equations<br />

r = r ( 1 2 3 ) (6.49)<br />

For the case of = − unknowns, any virtual displacement is <strong>in</strong>dependent of , therefore the<br />

only way for (644) to hold is for the term <strong>in</strong> brackets to vanish for each value of , thatis<br />

½ µ <br />

− ¾<br />

= <br />

(6.50)<br />

˙ <br />

where =1 2 3 These are the Lagrange equations for the m<strong>in</strong>imal set of <strong>in</strong>dependent generalized<br />

coord<strong>in</strong>ates.<br />

If all the generalized forces are conservative plus velocity <strong>in</strong>dependent, and are <strong>in</strong>cluded <strong>in</strong> the potential<br />

and <br />

=0,then(650) simplifies to<br />

½ µ <br />

− ¾<br />

=0 (6.51)<br />

˙ <br />

This is Euler’s differential equation, derived earlier us<strong>in</strong>g the calculus of variations. Thus d’Alembert’s<br />

Pr<strong>in</strong>ciple leads to a solution that m<strong>in</strong>imizes the action <strong>in</strong>tegral R 2<br />

1<br />

= 0 as stated by Hamilton’s<br />

Pr<strong>in</strong>ciple.<br />

6.5.3 Lagrange multipliers approach<br />

Equation (644) sums over all coord<strong>in</strong>ates for particles, provid<strong>in</strong>g equations of motion.<br />

constra<strong>in</strong>ts are holonomic they can be expressed by algebraic equations of constra<strong>in</strong>t<br />

If the <br />

( 1 2 )=0 (6.52)


6.5. CONSTRAINED SYSTEMS 139<br />

where =1 2 3 K<strong>in</strong>ematic constra<strong>in</strong>ts can be expressed <strong>in</strong> terms of the <strong>in</strong>f<strong>in</strong>itessimal displacements<br />

of the form<br />

X <br />

(q) + <br />

=0 (6.53)<br />

<br />

=1<br />

where =1 2 3, =1 2 3, and where the <br />

<br />

,and <br />

<br />

are functions of the generalized coord<strong>in</strong>ates<br />

, described by the vector q that are derived from the equations of constra<strong>in</strong>t. As discussed <strong>in</strong> chapter 57,<br />

if (653) represents the total differential of a function, then it can be <strong>in</strong>tegrated to give a holonomic relation<br />

of the form of equation (652). However, if (653) is not the total differential, then it can be <strong>in</strong>tegrated only<br />

after hav<strong>in</strong>g solved the full problem. If <br />

<br />

=0then the constra<strong>in</strong>t is scleronomic.<br />

The discussion of Lagrange multipliers <strong>in</strong> chapter 591, showed that, for virtual displacements <br />

the correlation of the generalized coord<strong>in</strong>ates, due to the constra<strong>in</strong>t forces, can be taken <strong>in</strong>to account by<br />

multiply<strong>in</strong>g (653) by unknown Lagrange multipliers and summ<strong>in</strong>g over all constra<strong>in</strong>ts. Generalized<br />

forces can be partitioned <strong>in</strong>to a Lagrange multiplier term plus a rema<strong>in</strong>der force. That is<br />

<br />

=<br />

X<br />

=1<br />

<br />

<br />

<br />

(q)+ <br />

(6.54)<br />

s<strong>in</strong>ce by def<strong>in</strong>ition =0for virtual displacements.<br />

Chapter 591 showed that holonomic forces of constra<strong>in</strong>t can be taken <strong>in</strong>to account by <strong>in</strong>troduc<strong>in</strong>g<br />

the Lagrange undeterm<strong>in</strong>ed multipliers approach, which is equivalent to def<strong>in</strong><strong>in</strong>g an extended Lagrangian<br />

0 (q ˙q λ) where<br />

0 (q ˙q λ) =(q ˙q)+<br />

X<br />

X<br />

=1 =1<br />

<br />

<br />

<br />

(q) (6.55)<br />

F<strong>in</strong>d<strong>in</strong>g the extremum for the extended Lagrangian 0 (q ˙q λ) us<strong>in</strong>g (647) gives<br />

"<br />

X ½ µ <br />

− ¾<br />

#<br />

X <br />

− (q) − <br />

=0 (6.56)<br />

˙ <br />

<br />

=1<br />

where <br />

is the rema<strong>in</strong><strong>in</strong>g part of the generalized force after subtract<strong>in</strong>g both the part of the force<br />

absorbed <strong>in</strong> the potential energy , which is buried <strong>in</strong> the Lagrangian , aswellastheholonomicconstra<strong>in</strong>t<br />

forces which are <strong>in</strong>cluded <strong>in</strong> the Lagrange multiplier terms P <br />

=1 <br />

<br />

(q). The Lagrange multipliers<br />

can be chosen arbitrarily <strong>in</strong> (656) Utiliz<strong>in</strong>g the free choice of the Lagrange multipliers allows them<br />

to be determ<strong>in</strong>ed <strong>in</strong> such a way that the coefficients of the first <strong>in</strong>f<strong>in</strong>itessimals, i.e. the square brackets<br />

vanish. Therefore the expression <strong>in</strong> the square bracket must vanish for each value of 1 ≤ ≤ . Thus it<br />

follows that<br />

½ µ <br />

− ¾ X <br />

− (q) − <br />

=0 (6.57)<br />

˙ <br />

=1<br />

when =1 2 Thus (656) reduces to a sum over the rema<strong>in</strong><strong>in</strong>g coord<strong>in</strong>ates between +1≤ ≤ <br />

"<br />

X ½ µ <br />

− ¾<br />

#<br />

X <br />

− (q) − <br />

=0 (6.58)<br />

˙ <br />

=+1<br />

=1<br />

In equation (658) the = − <strong>in</strong>f<strong>in</strong>itessimals canbechosenfreelys<strong>in</strong>cethe = − degrees<br />

of freedom are <strong>in</strong>dependent. Therefore the expression <strong>in</strong> the square bracket must vanish for each value of<br />

+1≤ ≤ . Thus it follows that<br />

½ µ <br />

− ¾ X <br />

− (q) − <br />

=0 (6.59)<br />

˙ <br />

=1<br />

where = +1+2 Comb<strong>in</strong><strong>in</strong>g equations (657) and (659) then gives the important general relation<br />

that for 1 ≤ ≤ <br />

½ µ <br />

− ¾ X <br />

= (q)+ <br />

(6.60)<br />

˙ <br />

=1


140 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

To summarize, the Lagrange multiplier approach (660) automatically solves the equations plus the<br />

holonomic equations of constra<strong>in</strong>t, which determ<strong>in</strong>es the + unknowns, that is, the coord<strong>in</strong>ates<br />

plus the forces of constra<strong>in</strong>t. The beauty of the Lagrange multipliers is that all variables, plus the <br />

constra<strong>in</strong>t forces, are found simultaneously by us<strong>in</strong>g the calculus of variations to determ<strong>in</strong>e the extremum<br />

for the expanded Lagrangian 0 (q ˙q λ).<br />

6.5.4 Generalized forces approach<br />

The two right-hand terms <strong>in</strong> (660) can be understood to be those forces act<strong>in</strong>g on the system that are<br />

not<br />

P<br />

absorbed <strong>in</strong>to the scalar potential component of the Lagrangian . The Lagrange multiplier terms<br />

<br />

=1 <br />

<br />

(q) account for the holonomic forces of constra<strong>in</strong>t that are not <strong>in</strong>cluded <strong>in</strong> the conservative<br />

potential or <strong>in</strong> the generalized forces <br />

. The generalized force<br />

<br />

=<br />

X<br />

<br />

F <br />

· r <br />

<br />

(617)<br />

is the sum of the components <strong>in</strong> the direction for all external forces that have not been taken <strong>in</strong>to account<br />

by the scalar potential or the Lagrange multipliers. Thus the non-conservative generalized force <br />

<br />

conta<strong>in</strong>s non-holonomic constra<strong>in</strong>t forces, <strong>in</strong>clud<strong>in</strong>g dissipative forces such as drag or friction, that are not<br />

<strong>in</strong>cluded <strong>in</strong> or used <strong>in</strong> the Lagrange multiplier terms to account for the holonomic constra<strong>in</strong>t forces.<br />

The concept of generalized forces is illustrated by the case of spherical coord<strong>in</strong>ate systems. The attached<br />

table gives the displacement elements ,(takenfromtable4) and the generalized force for the three<br />

coord<strong>in</strong>ates. Note that has the dimensions of force and has the units of energy. By contrast<br />

equation 630 gives that = and = which have the dimensions of torque. However, and<br />

bothhavethedimensionsofenergyasisrequired<strong>in</strong>equation630. This illustrates that the units used<br />

for generalized forces depend on the units of the correspond<strong>in</strong>g generalized coord<strong>in</strong>ate.<br />

Unit vectors · <br />

ˆ ˆr ˆr <br />

ˆθ ˆθ ˆθ <br />

ˆφ ˆφ s<strong>in</strong> ˆφ s<strong>in</strong> s<strong>in</strong> <br />

6.6 Apply<strong>in</strong>g the Euler-Lagrange equations to classical mechanics<br />

d’Alembert’s pr<strong>in</strong>ciple of virtual work has been used to derive the Euler-Lagrange equations, which also<br />

satisfy Hamilton’s Pr<strong>in</strong>ciple, and the Newtonian plausibility argument. These imply that the actual path<br />

taken<strong>in</strong>configuration space ( <br />

) is the one that m<strong>in</strong>imizes the action <strong>in</strong>tegral R 2<br />

1<br />

( <br />

; ) As a<br />

consequence, the Euler equations for the calculus of variations lead to the Lagrange equations of motion.<br />

½ µ <br />

− ¾ X <br />

= (q)+ <br />

<br />

(660)<br />

˙ <br />

=1<br />

for variables, with equations of constra<strong>in</strong>t. The generalized forces <br />

are not <strong>in</strong>cluded <strong>in</strong> the<br />

conservative, potential energy or the Lagrange multipliers approach for holonomic equations of constra<strong>in</strong>t. 2<br />

The follow<strong>in</strong>g is a logical procedure for apply<strong>in</strong>g the Euler-Lagrange equations to classical mechanics.<br />

1) Select a set of <strong>in</strong>dependent generalized coord<strong>in</strong>ates:<br />

Select an optimum set of <strong>in</strong>dependent generalized coord<strong>in</strong>ates as described <strong>in</strong> chapter 651. Use of generalized<br />

coord<strong>in</strong>ates is always advantageous s<strong>in</strong>ce they <strong>in</strong>corporate the constra<strong>in</strong>ts, and can reduce the number of<br />

unknowns, both of which simplify use of Lagrangian mechanics<br />

2 Euler’s differential equation is ubiquitous <strong>in</strong> Lagrangian mechanics. Thus, for brevity, it is convenient to def<strong>in</strong>e the concept<br />

of the Lagrange l<strong>in</strong>ear operator Λ as described <strong>in</strong> appendix 2<br />

Λ ≡ <br />

−<br />

<br />

˙ <br />

where Λ operates on the Lagrangian . Then Euler’s equations can be written compactly <strong>in</strong> the form Λ =0.


6.6. APPLYING THE EULER-LAGRANGE EQUATIONS TO CLASSICAL MECHANICS 141<br />

2) Partition of the active forces:<br />

Theactiveforcesshouldbepartitioned<strong>in</strong>tothefollow<strong>in</strong>gthreegroups:<br />

(i) Conservative one-body forces plus the velocity-dependent electromagnetic force which<br />

can be characterized by the scalar potential , that is absorbed <strong>in</strong>to the Lagrangian. The gravitational<br />

forces plus the velocity-dependent electromagnetic force can be absorbed <strong>in</strong>to the potential as discussed<br />

<strong>in</strong> chapter 610. This approach is by far the easiest way to account for such forces <strong>in</strong> Lagrangian mechanics.<br />

(ii) Holonomic constra<strong>in</strong>t forces provide algebraic relations that couple some of the generalized coord<strong>in</strong>ates.<br />

This coupl<strong>in</strong>g can be used either to reduce the number of generalized coord<strong>in</strong>ates, or to determ<strong>in</strong>e<br />

these holonomic constra<strong>in</strong>t forces us<strong>in</strong>g the Lagrange multiplier approach.<br />

(iii) Generalized forces provide a mechanism for <strong>in</strong>troduc<strong>in</strong>g non-conservative and non-holonomic<br />

constra<strong>in</strong>t forces <strong>in</strong>to Lagrangian mechanics. Typically general forces are used to <strong>in</strong>troduce dissipative<br />

forces.<br />

Typical systems can <strong>in</strong>volve a mixture of all three categories of active forces. For example, mechanical<br />

systems often <strong>in</strong>clude gravity, <strong>in</strong>troduced as a potential, holonomic constra<strong>in</strong>t forces are determ<strong>in</strong>ed us<strong>in</strong>g<br />

Lagrange multipliers, and dissipative forces are <strong>in</strong>cluded as generalized forces.<br />

3) M<strong>in</strong>imal set of generalized coord<strong>in</strong>ates:<br />

The ability to embed constra<strong>in</strong>t forces directly <strong>in</strong>to the generalized coord<strong>in</strong>ates is a tremendous advantage<br />

enjoyed by the Lagrangian and Hamiltonian variational approaches to classical mechanics. If the constra<strong>in</strong>t<br />

forces are not required, then choice of a m<strong>in</strong>imal set of generalized coord<strong>in</strong>ates significantly reduces the<br />

number of equations of motion that need to be solved .<br />

4) Derive the Lagrangian:<br />

The Lagrangian is derived <strong>in</strong> terms of the generalized coord<strong>in</strong>ates and <strong>in</strong>clud<strong>in</strong>g the conservative forces that<br />

are buried <strong>in</strong>to the scalar potential <br />

5) Derive the equations of motion:<br />

Equation (660) is solved to determ<strong>in</strong>e the generalized coord<strong>in</strong>ates, plus the Lagrange multipliers characteriz<strong>in</strong>g<br />

the holonomic constra<strong>in</strong>t forces, plus any generalized forces that were <strong>in</strong>cluded. The holonomic<br />

<br />

constra<strong>in</strong>t forces then are given by evaluat<strong>in</strong>g the <br />

(q) terms for the holonomic forces.<br />

In summary, Lagrangian mechanics is based on energies which are scalars <strong>in</strong> contrast to Newtonian<br />

mechanics which is based on vector forces and momentum. As a consequence, Lagrange mechanics allows<br />

use of any set of <strong>in</strong>dependent generalized coord<strong>in</strong>ates, which do not have to be orthogonal, and they can<br />

have very different units for different variables. The generalized coord<strong>in</strong>ates can <strong>in</strong>corporate the correlations<br />

<strong>in</strong>troduced by constra<strong>in</strong>t forces.<br />

The active forces are split <strong>in</strong>to the follow<strong>in</strong>g three categories;<br />

1. Velocity-<strong>in</strong>dependent conservative forces are taken <strong>in</strong>to account us<strong>in</strong>g scalar potentials .<br />

2. Holonomic constra<strong>in</strong>t forces can be determ<strong>in</strong>ed us<strong>in</strong>g Lagrange multipliers.<br />

3. Non-holonomic constra<strong>in</strong>ts require use of generalized forces <br />

.<br />

Use of the concept of scalar potentials is a trivial and powerful way to <strong>in</strong>corporate conservative forces <strong>in</strong><br />

Lagrangian mechanics. The Lagrange multipliers approach requires us<strong>in</strong>g the Euler-Lagrange equations for<br />

+ coord<strong>in</strong>ates but determ<strong>in</strong>es both holonomic constra<strong>in</strong>t forces and equations of motion simultaneously.<br />

Non-holonomic constra<strong>in</strong>ts and dissipative forces can be <strong>in</strong>corporated <strong>in</strong>to Lagrangian mechanics via use of<br />

generalized forces which broadens the scope of Lagrangian mechanics.<br />

Note that the equations of motion result<strong>in</strong>g from the Lagrange-Euler algebraic approach are the same<br />

equations of motion as obta<strong>in</strong>ed us<strong>in</strong>g Newtonian mechanics. However, the Lagrangian is a scalar which<br />

facilitates rotation <strong>in</strong>to the most convenient frame of reference. This can greatly simplify determ<strong>in</strong>ation of<br />

the equations of motion when constra<strong>in</strong>t forces apply. As discussed <strong>in</strong> chapter 17, the Lagrangian and the<br />

Hamiltonian variational approaches to mechanics are the only viable way to handle relativistic, statistical,<br />

and quantum mechanics.


142 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

6.7 Applications to unconstra<strong>in</strong>ed systems<br />

Although most dynamical systems <strong>in</strong>volve constra<strong>in</strong>ed motion, it is useful to consider examples of systems<br />

subject to conservative forces with no constra<strong>in</strong>ts. For no constra<strong>in</strong>ts, the Lagrange-Euler equations (660)<br />

simplify to Λ =0where =1 2 and the transformation to generalized coord<strong>in</strong>ates is of no consequence.<br />

6.1 Example: Motion of a free particle, U=0<br />

The Lagrangian <strong>in</strong> cartesian coord<strong>in</strong>ates is = 1 2 ( ˙2 + ˙ 2 + ˙ 2 ) Then<br />

<br />

˙<br />

<br />

˙<br />

<br />

˙<br />

<br />

<br />

= ˙<br />

= ˙<br />

= ˙<br />

= <br />

= <br />

=0<br />

Insert these <strong>in</strong> the Lagrange equation gives<br />

Thus<br />

Λ = <br />

˙ − <br />

= ˙ − 0=0<br />

<br />

= ˙ = <br />

= ˙ = <br />

= ˙ = <br />

That is, this shows that the l<strong>in</strong>ear momentum is conserved if is a constant, that is, no forces apply. Note<br />

that momentum conservation has been derived without any direct reference to forces.<br />

6.2 Example: Motion <strong>in</strong> a uniform gravitational field<br />

Consider the motion³ is <strong>in</strong> the − plane. The<br />

<br />

k<strong>in</strong>etic energy = 1 2 2 2´<br />

+ while the potential<br />

energy is = where ( =0)=0 Thus<br />

y<br />

= 1 2 ³ 2<br />

+<br />

<br />

<br />

2´ − <br />

Us<strong>in</strong>g the Lagrange equation for the coord<strong>in</strong>ate<br />

gives<br />

(x, y)<br />

g<br />

Λ = <br />

− <br />

= − 0=0<br />

Thus the horizontal momentum ˙ is conserved and<br />

<br />

=0 The coord<strong>in</strong>ate gives<br />

Λ = <br />

− <br />

= + =0<br />

Motion <strong>in</strong> a gravitational field<br />

x<br />

Thus the Lagrangian produces the same results as derived<br />

us<strong>in</strong>g Newton’s Laws of Motion.<br />

<br />

¨ =0<br />

= −


6.7. APPLICATIONS TO UNCONSTRAINED SYSTEMS 143<br />

The importance of select<strong>in</strong>g the most convenient generalized coord<strong>in</strong>ates is nicely illustrated by try<strong>in</strong>g to<br />

solve this problem us<strong>in</strong>g polar coord<strong>in</strong>ates where is radial distance and the elevation angle from the<br />

axis as shown <strong>in</strong> the adjacent figure. Then<br />

= 1 2 2 + 1 ³<br />

2 <br />

´2<br />

Thus<br />

= s<strong>in</strong> <br />

= 1 2 2 + 1 2 ³ ˙´2<br />

− s<strong>in</strong> <br />

Λ =0for the coord<strong>in</strong>ate<br />

˙ 2 − s<strong>in</strong> − ¨ =0<br />

Λ =0for the coord<strong>in</strong>ate<br />

− cos − 2 ˙ ˙ − 2¨ =0<br />

These equations written <strong>in</strong> polar coord<strong>in</strong>ates are more complicated than the result expressed <strong>in</strong> cartesian<br />

coord<strong>in</strong>ates. This is because the potential energy depends directly on the coord<strong>in</strong>ate, whereas it is a function<br />

of both This illustrates the freedom for us<strong>in</strong>g different generalized coord<strong>in</strong>ates, plus the importance of<br />

choos<strong>in</strong>g a sensible set of generalized coord<strong>in</strong>ates.<br />

6.3 Example: Central forces<br />

Consider a mass mov<strong>in</strong>g under the <strong>in</strong>fluence of a spherically-symmetric, conservative, attractive,<br />

<strong>in</strong>verse-square force. The potential then is<br />

= − <br />

It is natural to express the Lagrangian <strong>in</strong> spherical coord<strong>in</strong>ates for this system. That is,<br />

= 1 2 ˙2 + 1 ³<br />

2 ˙<br />

´2 1 +<br />

2 ( s<strong>in</strong> ˙) 2 + <br />

Λ =0for the coord<strong>in</strong>ate gives<br />

¨ − [˙ 2 +s<strong>in</strong> 2 ˙ 2 ]= 2<br />

where the s<strong>in</strong> 2 ˙ 2 term comes from the centripetal acceleration.<br />

Λ =0for the coord<strong>in</strong>ate gives<br />

<br />

³<br />

2 s<strong>in</strong> 2 <br />

<br />

˙<br />

´<br />

=0<br />

This implies that the derivative of the angular momentum about the axis, ˙ =0and thus = 2 s<strong>in</strong> 2 ˙<br />

is a constant of motion.<br />

Λ =0for the coord<strong>in</strong>ate gives<br />

<br />

(2 ˙) − 2 s<strong>in</strong> cos ˙ 2 =0<br />

That is,<br />

˙ = 2 s<strong>in</strong> cos ˙ 2 =<br />

2 cos <br />

2 2 s<strong>in</strong> 3 <br />

Note that is a constant of motion if =0and only the radial coord<strong>in</strong>ate is <strong>in</strong>fluenced by the radial form<br />

of the central potential.


144 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

6.8 Applications to systems <strong>in</strong>volv<strong>in</strong>g holonomic constra<strong>in</strong>ts<br />

The equations of motion that result from the Lagrange-Euler algebraic approach are the same as those given<br />

by Newtonian mechanics. The solution of these equations of motion can be obta<strong>in</strong>ed mathematically us<strong>in</strong>g<br />

the chosen <strong>in</strong>itial conditions. The follow<strong>in</strong>g simple example of a disk roll<strong>in</strong>g on an <strong>in</strong>cl<strong>in</strong>ed plane, is useful<br />

for compar<strong>in</strong>g the merits of the Newtonian method with Lagrange mechanics employ<strong>in</strong>g either m<strong>in</strong>imal<br />

generalized coord<strong>in</strong>ates, the Lagrange multipliers, or the generalized forces approaches.<br />

6.4 Example: Disk roll<strong>in</strong>g on an <strong>in</strong>cl<strong>in</strong>ed plane<br />

Consider a disk roll<strong>in</strong>g down an <strong>in</strong>cl<strong>in</strong>ed plane to compare<br />

the results obta<strong>in</strong>ed us<strong>in</strong>g Newton’s laws with the results obta<strong>in</strong>ed<br />

us<strong>in</strong>g Lagrange’s equations with either generalized coord<strong>in</strong>ates,<br />

Lagrange multipliers, or generalized forces. All these<br />

cases assume that the friction is sufficient to ensure that the<br />

roll<strong>in</strong>g equation of constra<strong>in</strong>t applies and that the disk has a<br />

radius and moment of <strong>in</strong>ertia of . Assume as generalized<br />

coord<strong>in</strong>ates, distance along the <strong>in</strong>cl<strong>in</strong>ed plane which is perpendicular<br />

to the normal constra<strong>in</strong>t force , and perpendicular<br />

to the <strong>in</strong>cl<strong>in</strong>ed plane , plus the roll<strong>in</strong>g angle . The constra<strong>in</strong>t<br />

for roll<strong>in</strong>g is holonomic<br />

− =0<br />

The frictional force is The constra<strong>in</strong>t that it rolls along the<br />

plane implies<br />

− =0<br />

y<br />

N<br />

F f<br />

mg<br />

Disk roll<strong>in</strong>g without slipp<strong>in</strong>g on an<br />

<strong>in</strong>cl<strong>in</strong>ed plane.<br />

a) Newton’s laws of motion<br />

Newton’s law for the components of the forces along the <strong>in</strong>cl<strong>in</strong>ed plane gives<br />

s<strong>in</strong> − = <br />

<br />

Perpendicular to the <strong>in</strong>cl<strong>in</strong>ed plane, Newton’s law gives<br />

The torque on the disk gives<br />

Assum<strong>in</strong>gthediscrollsgives<br />

then<br />

cos = <br />

= ¨<br />

<br />

= ¨<br />

= 2 ¨<br />

(a)<br />

(b)<br />

(c)<br />

Insert<strong>in</strong>g this <strong>in</strong>to equation (a) gives<br />

µ + 2 <br />

¨ − s<strong>in</strong> =0<br />

The moment of <strong>in</strong>ertia of a uniform solid circular disk is = 1 2 2<br />

Therefore<br />

¨ = 2 3 s<strong>in</strong> <br />

and the frictional force is<br />

= <br />

3 s<strong>in</strong> <br />

which is smaller than the gravitational force along the plane which is s<strong>in</strong>


6.8. APPLICATIONS TO SYSTEMS INVOLVING HOLONOMIC CONSTRAINTS 145<br />

b) Lagrange equations with a m<strong>in</strong>imal set of generalized coord<strong>in</strong>ates<br />

Us<strong>in</strong>g the generalized coord<strong>in</strong>ates def<strong>in</strong>ed above, the total k<strong>in</strong>etic energy is<br />

= 1 2 ˙2 + 1 2 ˙ 2<br />

The conservative gravitational force can be absorbed <strong>in</strong>to the potential energy<br />

= ( − )s<strong>in</strong><br />

Thus the Lagrangian is<br />

= 1 2 ˙2 + 1 2 ˙ 2 − ( − )s<strong>in</strong><br />

The holonomic equations of constra<strong>in</strong>t are<br />

1 = − =0<br />

2 = − =0<br />

A holonomic constra<strong>in</strong>t can be used to reduce the system to a s<strong>in</strong>gle generalized coord<strong>in</strong>ate plus generalized<br />

velocity ˙ Expressed <strong>in</strong> terms of this s<strong>in</strong>gle generalized coord<strong>in</strong>ate, the Lagrangian becomes<br />

= 1 µ + <br />

2 2 ˙ 2 − ( − )s<strong>in</strong><br />

The Lagrange equation Λ =0gives<br />

s<strong>in</strong> =<br />

µ + 2 <br />

<br />

<br />

Aga<strong>in</strong> if = 1 2 2 then<br />

¨ = 2 3 s<strong>in</strong> <br />

The solution for the coord<strong>in</strong>ate is trivial. This answer is identical to that obta<strong>in</strong>ed us<strong>in</strong>g Newton’s laws<br />

of motion. Note that no forces have been determ<strong>in</strong>ed us<strong>in</strong>g the s<strong>in</strong>gle generalized coord<strong>in</strong>ate.<br />

c) Lagrange equation with Lagrange multipliers<br />

Aga<strong>in</strong> the conservative gravitation force is absorbed <strong>in</strong>to the scalar potential while the holonomic constra<strong>in</strong>ts<br />

are taken <strong>in</strong>to account us<strong>in</strong>g Lagrange multipliers. Ignor<strong>in</strong>g the trivial dependence, the Lagrangian is given<br />

above to be<br />

= 1 2 ˙2 + 1 2 ˙ 2 − ( − )s<strong>in</strong><br />

The constra<strong>in</strong>t equations are<br />

The Lagrange equation for the coord<strong>in</strong>ate<br />

gives<br />

The Lagrange equation for the coord<strong>in</strong>ate<br />

1 = − =0<br />

2 = − =0<br />

<br />

˙ − <br />

= 1<br />

1<br />

+ 20<br />

<br />

<br />

¨ − s<strong>in</strong> = 1<br />

<br />

−<br />

˙ = 1<br />

1<br />

+ 20


146 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

which gives<br />

The constra<strong>in</strong>t can be written as<br />

Let = 1 2 2 and solve for and gives<br />

The frictional force is given by<br />

Also<br />

and the torque is<br />

d) Lagrange equation us<strong>in</strong>g a generalized force<br />

¨ = − 1 <br />

¨ = ¨<br />

<br />

1 = −¡ ¢ s<strong>in</strong> = − <br />

1+<br />

2<br />

3 s<strong>in</strong> <br />

<br />

= 1<br />

1<br />

= 1 = − <br />

3 s<strong>in</strong> <br />

¨ = s<strong>in</strong> + 1 = 2 s<strong>in</strong> <br />

3<br />

− 1 = = ¨<br />

Aga<strong>in</strong> the conservative gravitation force is absorbed <strong>in</strong>to the scalar potential while the holonomic constra<strong>in</strong>ts<br />

are taken <strong>in</strong>to account us<strong>in</strong>g generalized forces. Ignor<strong>in</strong>g the trivial dependence, the Lagrangian was given<br />

above to be<br />

= 1 2 ˙2 + 1 2 ˙ 2 − ( − )s<strong>in</strong><br />

The generalized forces (630) are<br />

= − <br />

= <br />

The Euler-Lagrange equations are:<br />

The Λ = Lagrange equation for the coord<strong>in</strong>ate<br />

¨ − s<strong>in</strong> = = − <br />

The Λ = Lagrange equation for the coord<strong>in</strong>ate<br />

¨ = = <br />

The constra<strong>in</strong>t equation gives that = and assum<strong>in</strong>g = 1 2 2 leads to the relation<br />

<br />

= = 2 ¨<br />

Substitute this equation <strong>in</strong>to the relation gives that<br />

¨ − s<strong>in</strong> = = − = 2 ¨<br />

Thus<br />

and<br />

¨ = 2 3 s<strong>in</strong> <br />

= − <br />

3 s<strong>in</strong> <br />

The four methods for handl<strong>in</strong>g the equations of constra<strong>in</strong>t all are equivalent and result <strong>in</strong> the same<br />

equations of motion. The scalar Lagrangian mechanics is able to calculate the vector forces act<strong>in</strong>g <strong>in</strong> a direct<br />

and simple way. The Newton’s law approach is more <strong>in</strong>tuitive for this simple case and the ease and power<br />

of the Lagrangian approach is not apparent for this simple system.<br />

The follow<strong>in</strong>g series of examples will gradually <strong>in</strong>crease <strong>in</strong> complexity, and will illustrate the power,<br />

elegance, plus superiority of the Lagrangian approach compared with the Newtonian approach.


6.8. APPLICATIONS TO SYSTEMS INVOLVING HOLONOMIC CONSTRAINTS 147<br />

6.5 Example: Two connected masses on frictionless <strong>in</strong>cl<strong>in</strong>ed planes<br />

Consider the system shown <strong>in</strong> the figure. This is<br />

a problem that has five constra<strong>in</strong>ts that will be solved<br />

us<strong>in</strong>g the method of generalized coord<strong>in</strong>ates. The obvious<br />

generalized coord<strong>in</strong>ates are 1 and 2 which are<br />

perpendicular to the normal constra<strong>in</strong>t forces on the<br />

<strong>in</strong>cl<strong>in</strong>ed planes. Another holonomic constra<strong>in</strong>t is that<br />

the length of the rope connect<strong>in</strong>g the masses is assumed<br />

to be constant. Thus the equation of constra<strong>in</strong>t is that<br />

1 + 2 − =0<br />

1 2<br />

x 1<br />

x 2<br />

m 1 m 2<br />

The other four constra<strong>in</strong>ts ensure that the two masses<br />

slide directly down the <strong>in</strong>cl<strong>in</strong>ed planes <strong>in</strong> the plane<br />

shown. This is assumed implicitly by us<strong>in</strong>g only the<br />

variables, 1 and 2 Let us chose 1 as the primary<br />

generalized coord<strong>in</strong>ate, thus<br />

Two connected masses on frictionless <strong>in</strong>cl<strong>in</strong>ed<br />

planes<br />

2 = − 1<br />

1 = 1 s<strong>in</strong> 1<br />

2 = ( − 1 )s<strong>in</strong> 2<br />

The conservative gravitational force is absorbed <strong>in</strong>to the potential energy given by<br />

= − 1 1 s<strong>in</strong> 1 − 2 ( − 1 )s<strong>in</strong> 2<br />

S<strong>in</strong>ce 1 = − 2 the k<strong>in</strong>etic energy is given by<br />

The Lagrangian then gives that<br />

= 1 2 1 ˙ 2 1 + 1 2 2 ˙ 2 2 = 1 2 ( 1 + 2 ) ˙ 2 1<br />

= 1 2 ( 1 + 2 ) ˙ 2 1 + 1 1 s<strong>in</strong> 1 + 2 ( − 1 )s<strong>in</strong> 2<br />

x<br />

Therefore<br />

<br />

= ( 1 + 2 ) ˙ 1<br />

˙ 1<br />

<br />

= ( 1 s<strong>in</strong> 1 − 2 s<strong>in</strong> 2 )<br />

1<br />

Thus<br />

Λ 1 = <br />

− =0=( 1 + 2 )¨ 1 − ( 1 s<strong>in</strong> 1 − 2 s<strong>in</strong> 2 )<br />

˙ 1 1<br />

Note that the system acts as though the <strong>in</strong>ertial mass is ( 1 + 2 )<br />

while the driv<strong>in</strong>g force comes from the difference of the forces. The<br />

acceleration is zero if<br />

1 s<strong>in</strong> 1 = 2 s<strong>in</strong> 2<br />

A special case of this is the Atwood’s mach<strong>in</strong>e with a massless<br />

pulley shown <strong>in</strong> the adjacent figure. For this case 1 = 2 =90 <br />

Thus<br />

( 1 + 2 )¨ 1 = ( 1 − 2 )<br />

Note that this problem has been solved without any reference to the<br />

force <strong>in</strong> the rope or the normal constra<strong>in</strong>t forces on the <strong>in</strong>cl<strong>in</strong>ed planes.<br />

x<br />

1 2<br />

m<br />

1<br />

m<br />

Atwoods mach<strong>in</strong>e<br />

2


148 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

6.6 Example: Two blocks connected by a frictionless<br />

bar<br />

Two identical masses are connected by a massless<br />

rigid bar of length , and they are constra<strong>in</strong>ed to move<br />

<strong>in</strong> two frictionless slides, one vertical and the other horizontal<br />

as shown <strong>in</strong> the adjacent figure. Assume that the<br />

conservative gravitational force acts along the negative <br />

axis and is <strong>in</strong>corporated <strong>in</strong>to the scalar potential . The<br />

generalized coord<strong>in</strong>ate can be chosen to be the angle <br />

correspond<strong>in</strong>g to a s<strong>in</strong>gle degree of freedom. The relative<br />

cartesian coord<strong>in</strong>ates of the blocks are given by<br />

y<br />

l<br />

= cos <br />

= s<strong>in</strong> <br />

Thus<br />

˙ = −(s<strong>in</strong> ) ˙<br />

˙ = (cos ) ˙<br />

This constra<strong>in</strong>t, that is absorbed <strong>in</strong>to the generalized coord<strong>in</strong>ate,<br />

is holonomic, scleronomic, and conservative.<br />

The k<strong>in</strong>etic energy is given by<br />

Two frictionless masses that are connected by a<br />

bar and are constra<strong>in</strong>ed to slide <strong>in</strong> vertical and<br />

horizontal channels.<br />

x<br />

The gravitational potential energy is given by<br />

= 1 2 ¡ 2 (s<strong>in</strong> ) 2 ˙ 2 + 2 (cos ) 2 ˙ 2¢ = 1 2 2 ˙ 2<br />

= = s<strong>in</strong> <br />

Thus the Lagrangian is<br />

= 1 2 2 ˙ 2 − s<strong>in</strong> <br />

Us<strong>in</strong>g the Lagrange operator equation Λ =0gives<br />

2¨ + cos = 0<br />

¨ + cos = 0<br />

Multiply by ˙ yields<br />

¨ ˙ + <br />

˙ cos =0<br />

This can be <strong>in</strong>tegrated to give<br />

where is a constant. That is<br />

1<br />

2 ˙2 + s<strong>in</strong> = <br />

r<br />

˙ = 2<br />

³ − s<strong>in</strong> ´<br />

<br />

Separation of the variable gives<br />

<br />

= q2 ¡ − <br />

s<strong>in</strong> ¢<br />

Integration of this gives<br />

− 0 =<br />

Z <br />

0<br />

<br />

q2 ¡ − <br />

s<strong>in</strong> ¢<br />

The constants and 0 are determ<strong>in</strong>ed from the given <strong>in</strong>itial conditions.


6.8. APPLICATIONS TO SYSTEMS INVOLVING HOLONOMIC CONSTRAINTS 149<br />

6.7 Example: Block slid<strong>in</strong>g on a movable frictionless <strong>in</strong>cl<strong>in</strong>ed plane<br />

Consider a block of mass free to slide on a smooth<br />

frictionless <strong>in</strong>cl<strong>in</strong>ed plane of mass that is free to slide<br />

horizontally as shown <strong>in</strong> the adjacent figure. The six degrees<br />

of freedom can be reduced to two <strong>in</strong>dependent generalized<br />

coord<strong>in</strong>ates s<strong>in</strong>ce the <strong>in</strong>cl<strong>in</strong>ed plane and mass <br />

x’<br />

are conf<strong>in</strong>edtoslidealongspecific non-orthogonal directions.<br />

Choose as the coord<strong>in</strong>ate for movement of the<br />

<strong>in</strong>cl<strong>in</strong>ed plane <strong>in</strong> the horizontal ˆ direction and 0 the<br />

m<br />

position of the block with respect to the surface of the<br />

M<br />

<strong>in</strong>cl<strong>in</strong>ed plane <strong>in</strong> the ê direction which is <strong>in</strong>cl<strong>in</strong>ed downward<br />

at an angle . Thus the velocity of the <strong>in</strong>cl<strong>in</strong>ed<br />

x<br />

plane is<br />

V = ˆ ˙<br />

A block slid<strong>in</strong>g on a frictionless movable <strong>in</strong>cl<strong>in</strong>ed<br />

while the velocity of the small block on the <strong>in</strong>cl<strong>in</strong>ed plane<br />

plane.<br />

is<br />

v = ˆ ˙ + ê ˙ 0<br />

The k<strong>in</strong>etic energy is given by<br />

= 1 2 V · V+1 2 v · v = 1 2 ˙2 + 1 2 [ ˙2 + ˙ 02 +2˙ ˙ 0 cos ]<br />

The conservative gravitational force is absorbed <strong>in</strong>to the scalar potential energy which depends only on the<br />

vertical position of the block and is taken to be zero at the top of the wedge.<br />

= − 0 s<strong>in</strong> <br />

Thus the Lagrangian is<br />

= 1 2 ˙2 + 1 2 [ ˙2 + ˙ 02 +2˙ ˙ 0 cos ]+ 0 s<strong>in</strong> <br />

Consider the Lagrange-Euler equation for the coord<strong>in</strong>ate, Λ =0which gives<br />

<br />

[( ˙ + ˙0 cos )+ ˙] =0<br />

()<br />

which states that [( ˙ + ˙ 0 cos )+ ˙] is a constant of motion. This constant of motion is just the total<br />

l<strong>in</strong>ear momentum of the complete system <strong>in</strong> the direction. That is, conservation of the l<strong>in</strong>ear momentum<br />

is satisfied automatically by the Lagrangian approach. The Newtonian approach also predicts conservation of<br />

the l<strong>in</strong>ear momentum s<strong>in</strong>ce there are no external horizontal forces,<br />

Consider the Lagrangian equation for the 0 coord<strong>in</strong>ate Λ 0 =0which gives<br />

<br />

[ ˙0 + ˙ cos ] = s<strong>in</strong> <br />

Perform both of the time derivatives for equations and give<br />

Solv<strong>in</strong>g for ¨ and ¨ 0 gives<br />

[¨ +¨ 0 cos ]+ ¨ = 0<br />

¨ 0 +¨ cos = s<strong>in</strong> <br />

¨ =<br />

− s<strong>in</strong> cos <br />

( + ) − cos 2 <br />

and.<br />

¨ 0 s<strong>in</strong> <br />

=<br />

1 − cos 2 ( + )<br />

This example illustrates the flexibility of be<strong>in</strong>g able to use non-orthogonal displacement vectors to specify the<br />

scalar Lagrangian energy. Newtonian mechanics wouldrequiremorethoughttosolvethisproblem.<br />

()


150 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

6.8 Example: Sphere roll<strong>in</strong>g without slipp<strong>in</strong>g down an <strong>in</strong>cl<strong>in</strong>ed plane on a<br />

frictionless floor.<br />

A sphere of mass and radius rolls, without slipp<strong>in</strong>g, down an <strong>in</strong>cl<strong>in</strong>ed plane, of mass sitt<strong>in</strong>g on a<br />

frictionless horizontal floor as shown <strong>in</strong> the adjacent figure. The velocity of the roll<strong>in</strong>g sphere has horizontal<br />

and vertical components of<br />

= ˙ + ˙ cos <br />

= −˙ s<strong>in</strong> <br />

Assume <strong>in</strong>itial conditions are =0 =0=0 =0 = ˙ = ˙ =0 Choose the <strong>in</strong>dependent coord<strong>in</strong>ates<br />

and as generalized coord<strong>in</strong>ates plus the holonomic constra<strong>in</strong>t = . Then the Lagrangian is<br />

= 2 ˙2 + 2<br />

Lagrange’s equations Λ =0and Λ =0,give<br />

h<br />

˙ 2 + 2 ˙2 +2 ˙˙ cos <br />

i<br />

+ 5 2 ˙2 − ( − s<strong>in</strong> )<br />

( + )¨ + ¨ cos = 0<br />

y<br />

Elim<strong>in</strong>at<strong>in</strong>g ¨ gives<br />

¨ cos + 7 5 ¨ − s<strong>in</strong> = 0<br />

µ 7<br />

5 − cos2 <br />

+ <br />

<br />

¨ = <br />

s<strong>in</strong> <br />

<br />

Integrate this equation assum<strong>in</strong>g the <strong>in</strong>itial conditions,<br />

results <strong>in</strong><br />

=<br />

5( + )s<strong>in</strong><br />

2[7( + ) − 5 cos 2 ] 2<br />

Thus<br />

cos <br />

= −<br />

+ = 5 s<strong>in</strong> (2)<br />

4[7( + ) − 5 cos 2 ] 2<br />

x<br />

y<br />

.<br />

Solid sphere roll<strong>in</strong>g without slipp<strong>in</strong>g on an<br />

<strong>in</strong>cl<strong>in</strong>ed plane on a frictionless horizontal floor.<br />

x<br />

Note that these equations predict conservation of l<strong>in</strong>ear<br />

momentum for the block plus sphere.<br />

6.9 Example: Mass slid<strong>in</strong>g on a rotat<strong>in</strong>g straight frictionless rod.<br />

Consider a mass slid<strong>in</strong>g on a frictionless rod that<br />

rotates about one end of the rod with an angular velocity<br />

<br />

. Choose and to be generalized coord<strong>in</strong>ates. Then<br />

the k<strong>in</strong>etic energy is given by<br />

= 1 2 ˙2 + 1 2 2 ˙2<br />

.<br />

m<br />

and potential energy<br />

=0<br />

The Lagrange equation for gives<br />

Λ = <br />

−<br />

˙ = (2 ˙) =0<br />

Thus the angular momentum is constant<br />

Mass slid<strong>in</strong>g on a rotat<strong>in</strong>g straight frictionless<br />

rod.<br />

2 ˙ = constant =


6.8. APPLICATIONS TO SYSTEMS INVOLVING HOLONOMIC CONSTRAINTS 151<br />

The Lagrange equation for gives<br />

Λ = <br />

˙ − <br />

= ¨ − ˙ 2 =0<br />

The equation states that the angular momentum is conserved for this case which is what we expect s<strong>in</strong>ce<br />

there are no external torques act<strong>in</strong>g on the system. The equation states that the centrifugal acceleration is<br />

¨ = 2 These equations of motion were derived without reference to the forces between the rod and mass.<br />

6.10 Example: Spherical pendulum<br />

The spherical pendulum is a classic holonomic<br />

problem <strong>in</strong> mechanics that <strong>in</strong>volves rotation plus oscillation<br />

where the pendulum is free to sw<strong>in</strong>g <strong>in</strong> any<br />

direction. This also applies to a particle constra<strong>in</strong>ed<br />

to slide <strong>in</strong> a smooth frictionless spherical bowl under<br />

gravity, such as a bar of soap <strong>in</strong> a wet hemispherical<br />

s<strong>in</strong>k. Consider the equation of motion of the spherical<br />

pendulum of mass and length shown <strong>in</strong> the<br />

adjacent figure. The most convenient generalized coord<strong>in</strong>ates<br />

are with orig<strong>in</strong> at the fulcrum, s<strong>in</strong>ce<br />

the length is constra<strong>in</strong>ed to be = The k<strong>in</strong>etic<br />

energy is<br />

= 1 1<br />

2 2 ˙2 +<br />

2 2 s<strong>in</strong> 2 ˙ 2<br />

The potential energy<br />

g<br />

m<br />

giv<strong>in</strong>g that<br />

= − cos <br />

= 1 1<br />

2 2 ˙2 +<br />

2 2 s<strong>in</strong> 2 ˙ 2 + cos <br />

The Lagrange equation for <br />

Λ = <br />

which gives<br />

The Lagrange equation for <br />

<br />

−<br />

˙ =0<br />

2¨ = <br />

2 ˙2 s<strong>in</strong> cos − s<strong>in</strong> <br />

Λ = <br />

−<br />

˙ = [2 s<strong>in</strong> 2 ˙] =0<br />

Spherical pendulum<br />

which gives<br />

2 s<strong>in</strong> 2 ˙ = = constant<br />

This is just the angular momentum for the pendulum rotat<strong>in</strong>g <strong>in</strong> the direction. Automatically the<br />

Lagrange approach shows that the angular momentum is a conserved quantity. This is what is expected<br />

from Newton’s Laws of Motion s<strong>in</strong>ce there are no external torques applied about this vertical axis.<br />

The equation of motion for can be simplified to<br />

¨ + s<strong>in</strong> −<br />

2 cos <br />

2 4 s<strong>in</strong> 3 =0<br />

There are many possible solutions depend<strong>in</strong>g on the <strong>in</strong>itial conditions. The pendulum can just oscillate<br />

<strong>in</strong> the direction, or rotate <strong>in</strong> the direction or some comb<strong>in</strong>ation of these. Note that if is zero, then<br />

the equation reduces to the simple harmonic pendulum, while the other extreme is when ¨ =0for which the<br />

motion is that of a conical pendulum that rotates at a constant angle 0 to the vertical axis.


152 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

6.11 Example: Spr<strong>in</strong>g plane pendulum<br />

Amass is suspended by a spr<strong>in</strong>g with spr<strong>in</strong>g constant <strong>in</strong> the gravitational field. Besides the longitud<strong>in</strong>al<br />

spr<strong>in</strong>g vibration, the spr<strong>in</strong>g performs a plane pendulum motion <strong>in</strong> the vertical plane, as illustrated <strong>in</strong><br />

the adjacent figure. F<strong>in</strong>d the Lagrangian, the equations of motion, and force <strong>in</strong> the spr<strong>in</strong>g.<br />

The system is holonomic, conservative, and scleronomic. Introduce plane polar coord<strong>in</strong>ates with radial<br />

length and polar angle as generalized coord<strong>in</strong>ates. The generalized coord<strong>in</strong>ates are related to the cartesian<br />

coord<strong>in</strong>ates by<br />

= cos <br />

= s<strong>in</strong> <br />

Therefore the velocities are given by<br />

˙ = ˙ cos + ˙ s<strong>in</strong> <br />

˙ = ˙ s<strong>in</strong> − ˙ cos <br />

r<br />

The k<strong>in</strong>etic energy is given by<br />

= 1 2 ¡ ˙ 2 + ˙ 2¢ = 1 2 ³<br />

˙ 2 + 2 ˙2´<br />

The gravitational plus spr<strong>in</strong>g potential energies both can be absorbed<br />

<strong>in</strong>to the potential .<br />

= − cos + 2 ( − 0) 2<br />

where 0 denotes the rest length of the spr<strong>in</strong>g. The Lagrangian thus equals<br />

= 1 ³<br />

2 ˙ 2 + ˙2´<br />

2 + cos − 2 ( − 0) 2<br />

y<br />

Spr<strong>in</strong>g pendulum hav<strong>in</strong>g spr<strong>in</strong>g<br />

constant and oscillat<strong>in</strong>g <strong>in</strong> a<br />

vertical plane.<br />

m<br />

For the polar angle , the Lagrange equation Λ =0gives<br />

<br />

³ ´<br />

2 ˙ = − s<strong>in</strong> <br />

<br />

The angular momentum = 2 ˙, thus the equation of motion can be written as<br />

Alternatively, evaluat<strong>in</strong>g <br />

³ ´<br />

2 ˙ gives<br />

˙ = − s<strong>in</strong> <br />

2¨ = − s<strong>in</strong> − 2 ˙ ˙<br />

The last term <strong>in</strong> the right-hand side is the Coriolis force caused by the time variation of the pendulum length.<br />

For the radial distance theLagrangeequationΛ =0gives<br />

¨ = ˙ 2 + cos − ( − 0 )<br />

This equation just equals the tension <strong>in</strong> the spr<strong>in</strong>g, i.e. = ¨. The first term on the right-hand side<br />

represents the centrifugal radial acceleration, the second term is the component of the gravitational force,<br />

and the third term represents Hooke’s Law for the spr<strong>in</strong>g. For small amplitudes of the motion appears as<br />

a superposition of harmonic oscillations <strong>in</strong> the plane.<br />

In this example the orthogonal coord<strong>in</strong>ate approach used gave the tension <strong>in</strong> the spr<strong>in</strong>g thus it is unnecessary<br />

to repeat this us<strong>in</strong>g the Lagrange multiplier approach.


6.8. APPLICATIONS TO SYSTEMS INVOLVING HOLONOMIC CONSTRAINTS 153<br />

6.12 Example: The yo-yo<br />

Considerayo-yocompris<strong>in</strong>gadiscthathasastr<strong>in</strong>gwrappedarounditwithoneendattachedtoafixed<br />

support. The disc is allowed to fall with the str<strong>in</strong>g unw<strong>in</strong>d<strong>in</strong>g as it falls as illustrated <strong>in</strong> the adjacent figure.<br />

Derive the equations of motion and the forces of constra<strong>in</strong>t via use of Lagrange multipliers. Use and as<br />

<strong>in</strong>dependent generalized coord<strong>in</strong>ates.<br />

The k<strong>in</strong>etic energy of the fall<strong>in</strong>g yo-yo is given by<br />

= 1 2 ˙2 + 1 2 ˙ 2 = 1 2 ˙2 + 1 4 2 ˙2<br />

where is the mass of the disc, the radius, and =<br />

1<br />

2 2 is the moment of <strong>in</strong>ertia of the disc about its central<br />

axis. The potential energy of the disc is<br />

= −<br />

y<br />

Thus the Lagrangian is<br />

= 1 2 ˙2 + 1 4 2 ˙2 + <br />

The one equation of constra<strong>in</strong>t is holonomic<br />

( ) = − =0<br />

The two Lagrange equations are<br />

<br />

− <br />

0 + =0<br />

<br />

− <br />

0 + <br />

=0<br />

The yo-yo comprises a fall<strong>in</strong>g disc unroll<strong>in</strong>g<br />

from a str<strong>in</strong>g attached to the disc at one end<br />

and a fixed support at the other end.<br />

with only one Lagrange multiplier . Evaluat<strong>in</strong>g these two Euler-Lagrange equations leads to two equations<br />

of motion<br />

Differentiat<strong>in</strong>g the equation of constra<strong>in</strong>t gives<br />

− ¨ + = 0<br />

− 1 2 2¨ − = 0<br />

¨ = ¨ <br />

Insert<strong>in</strong>g this <strong>in</strong>to the second equation and solv<strong>in</strong>g the two equations gives<br />

Insert<strong>in</strong>g <strong>in</strong>to the two equations of motion gives<br />

Thegeneralizedforceofconstra<strong>in</strong>t<br />

and the constra<strong>in</strong>t torque is<br />

= − 1 3 <br />

¨ = 2 3 <br />

¨ = 2 <br />

3 <br />

= <br />

= −1 3 <br />

= <br />

= 1 3 <br />

Thus the str<strong>in</strong>g reduces the acceleration of the disc <strong>in</strong> the gravitational field by a factor of<br />

1<br />

3 .


154 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

6.13 Example: Mass constra<strong>in</strong>ed to move on the <strong>in</strong>side of a frictionless paraboloid<br />

Amass moves on the frictionless <strong>in</strong>ner surface of a paraboloid<br />

2 + 2 = 2 = <br />

z<br />

with a gravitational potential energy of = <br />

This system is holonomic, scleronomic, and conservative. Choose<br />

cyl<strong>in</strong>drical coord<strong>in</strong>ates with respect to the vertical axis of the<br />

paraboloid to be the generalized coord<strong>in</strong>ates.<br />

The Lagrangian is<br />

= 1 2 ³<br />

˙ 2 + 2 ˙2 + ˙<br />

2´ − <br />

g<br />

z<br />

The equation of constra<strong>in</strong>t is<br />

( ) = 2 − =0<br />

The Lagrange multiplier approach will be used to determ<strong>in</strong>e the forces<br />

of constra<strong>in</strong>t.<br />

For Λ = <br />

<br />

x<br />

Mass constra<strong>in</strong>ed to slide on the<br />

<strong>in</strong>side of a frictionless paraboloid.<br />

r<br />

y<br />

<br />

˙ − <br />

<br />

2´<br />

<br />

³¨ − ˙<br />

= 1 2 (a)<br />

= 1 2<br />

For Λ = <br />

<br />

³ ´<br />

2 ˙ = ˙ =0 (b)<br />

<br />

<br />

Thus the angular momentum is conserved, that is, it is a constant of motion.<br />

For Λ = <br />

<br />

¨ = − − 1 <br />

and the time differential of the constra<strong>in</strong>t equation is<br />

(c)<br />

2˙ − ˙ =0<br />

(d)<br />

The above four equations of motion can be used to determ<strong>in</strong>e 1 <br />

The radius of the circle at the <strong>in</strong>tersection of the plane = with the paraboloid 2 = is given by<br />

0 = √ For a constant height = , then ¨ =0and equation (c) reduces to<br />

1 = − <br />

<br />

Therefore the constra<strong>in</strong>t force is given by<br />

= 1<br />

( )<br />

<br />

= − <br />

2<br />

Assum<strong>in</strong>g that ¨ =0 then equation (a) for ˙ = and = 0 gives<br />

¡ 0 − 0 2¢ = 1 2 0 = − <br />

2 0 = <br />

That is, the constra<strong>in</strong>t force equals<br />

= − 0 2<br />

which is the usual centripetal force. These relations also give that the <strong>in</strong>itial angular velocity required for<br />

such a stable trajectory with height is<br />

r<br />

2<br />

˙ = =


6.8. APPLICATIONS TO SYSTEMS INVOLVING HOLONOMIC CONSTRAINTS 155<br />

6.14 Example: Mass on a frictionless plane connected to a plane pendulum<br />

Two masses 1 and 2 areconnectedbyastr<strong>in</strong>gof<br />

length . Mass 1 is on a horizontal frictionless table<br />

and it is assumed that mass 2 moves <strong>in</strong> a vertical plane.<br />

This is another problem <strong>in</strong>volv<strong>in</strong>g holonomic constra<strong>in</strong>ed<br />

motion. The constra<strong>in</strong>ts are:<br />

1) 1 moves <strong>in</strong> the horizontal plane<br />

2) 2 moves <strong>in</strong> the vertical plane<br />

3) + = Therefore ˙ = − ˙<br />

There are 6−3 =3rema<strong>in</strong><strong>in</strong>g degrees of freedom after<br />

tak<strong>in</strong>g the constra<strong>in</strong>ts <strong>in</strong>to account. Choose as a set of<br />

generalized coord<strong>in</strong>ates, and In terms of these three<br />

generalized coord<strong>in</strong>ates, the k<strong>in</strong>etic energy is<br />

= 1 2 1<br />

= 1 2 1<br />

³<br />

˙ 2 + 2 ˙2´ + 1 2 2<br />

³<br />

˙ 2 +( − ) 2 ˙2´ + 1 2 2<br />

³<br />

˙ 2 + 2 ˙2´<br />

µ<br />

˙ 2 + 2 <br />

· 2 <br />

The potential energy <strong>in</strong> terms of the generalized coord<strong>in</strong>ates<br />

relative to the horizontal plane, is<br />

=0− 2 cos <br />

Therefore the Lagrangian equals<br />

The differentials are<br />

= 1 2 1<br />

m<br />

2<br />

r<br />

Mass 2 hang<strong>in</strong>g from a rope that is connected<br />

to 1 which slides on a frictionless plane.<br />

³<br />

˙ 2 +( − ) ˙2´ 2 + 1 ³<br />

2 2 ˙ 2 + ˙2´ 2 + 2 cos <br />

<br />

<br />

= −( − ) ˙ 2 + 2 ˙ 2 + cos <br />

<br />

˙<br />

= ( 1 + 2 ) ˙<br />

<br />

<br />

= − s<strong>in</strong> <br />

<br />

˙<br />

= 2 2 ˙<br />

<br />

= 0<br />

<br />

˙ = 1 ( − ) 2 ˙<br />

Thus the three Lagrange equations are<br />

Λ = ( 1 + 2 )¨ + 1 ( − ) ˙ 2 − 2 ˙ 2 − 2 cos =0<br />

Λ = h i<br />

2 2 ˙ + 2 s<strong>in</strong> =0<br />

<br />

s<br />

m 1<br />

that is<br />

2 2 ˙ ˙ + 2 2¨ + 2 s<strong>in</strong> =0<br />

Λ = h<br />

i<br />

1 ( − ) 2 ˙ =0<br />

<br />

This last equation is a statement of the conservation of angular momentum. These three differential equations<br />

of motion can be solved for known <strong>in</strong>itial conditions.


156 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

6.15 Example: Two connected masses constra<strong>in</strong>ed to slide along a mov<strong>in</strong>g rod<br />

Consider two identical masses constra<strong>in</strong>edtomove<br />

along the axis of a th<strong>in</strong> straight rod, of mass and length<br />

which is free to both translate and rotate. Two identical<br />

spr<strong>in</strong>gs l<strong>in</strong>k the two masses to the central po<strong>in</strong>t of the<br />

rod. Consider only motions of the system for which the<br />

extended lengths of the two spr<strong>in</strong>gs are equal and opposite<br />

such that the two masses always are equal distances from<br />

the center of the rod keep<strong>in</strong>g the center of mass at the<br />

center of the rod. F<strong>in</strong>d the equations of motion for this<br />

system.<br />

Use a fixed cartesian coord<strong>in</strong>ate system ( ) and<br />

a mov<strong>in</strong>g frame with the orig<strong>in</strong> at the center of the<br />

rod with its cartesian coord<strong>in</strong>ates ( 1 1 1 ) be<strong>in</strong>g parallel<br />

to the fixed coord<strong>in</strong>ate frame as shown <strong>in</strong> the figure. Let<br />

( ) be the spherical coord<strong>in</strong>ates of a po<strong>in</strong>t referr<strong>in</strong>g to<br />

the center of the mov<strong>in</strong>g ( 1 1 1 ) frame as shown <strong>in</strong> the<br />

figure. Then the two masses have spherical coord<strong>in</strong>ates<br />

( ) and (− ) <strong>in</strong> the mov<strong>in</strong>g-rod fixed frame. The<br />

frictionless constra<strong>in</strong>ts are holonomic.<br />

x<br />

z<br />

y<br />

x<br />

1<br />

r<br />

Two identical masses constra<strong>in</strong>ed to slide on<br />

amov<strong>in</strong>grodofmass The masses are<br />

attached to the center of the rod by identical<br />

spr<strong>in</strong>gs each hav<strong>in</strong>g a spr<strong>in</strong>g constant .<br />

The k<strong>in</strong>etic energy of the system is equal to the k<strong>in</strong>etic energy for all the mass concentrated at the center<br />

of mass plus the k<strong>in</strong>etic energy about the center of mass. S<strong>in</strong>ce is the center of mass then the k<strong>in</strong>etic<br />

energy can be separated <strong>in</strong>to three terms<br />

= + <br />

<br />

+ <br />

<br />

Note that s<strong>in</strong>ce the k<strong>in</strong>etic energy is a scalar quantity it is rotational <strong>in</strong>variant and thus can be evaluated <strong>in</strong><br />

any rotated frame. Thus the k<strong>in</strong>etic energy of the center of mass is<br />

= 1 2 ( +2)( ˙2 + ˙ 2 + ˙ 2 )<br />

The rotational k<strong>in</strong>etic energy of the two masses <strong>in</strong> the center of mass frame is<br />

<br />

= ( ˙ 2 + 2 ˙2 + 2 ˙ 2 s<strong>in</strong> 2 )<br />

The rotational k<strong>in</strong>etic energy of the rod <br />

is a scalar and thus can be evaluated <strong>in</strong> any rotated frame of<br />

reference fixed with respect to the pr<strong>in</strong>cipal axis system of the rod. The angular velocity of the rod about <br />

resolved along its pr<strong>in</strong>cipal axes is given by<br />

¯ = ˙ cos ê − ˙ s<strong>in</strong> ê − ˙ê <br />

The correspond<strong>in</strong>g moments of <strong>in</strong>ertia of the uniform <strong>in</strong>f<strong>in</strong>itesimally-th<strong>in</strong> rod are =0 = 1 12 2 =<br />

1<br />

12 2 . Hence the rotational k<strong>in</strong>etic energy of the rod is<br />

= 1 2 ( 2 + 2 + 2 )= 1<br />

24 2 (˙ 2 + ˙ 2 s<strong>in</strong> 2 )<br />

Theonlypotentialenergyisduetothetwoextendedspr<strong>in</strong>gswhichareassumedtohavethesamelength<br />

where 0 is the unstretched length.<br />

=2· 1<br />

2 ( − 0) 2 = ( − 0 ) 2<br />

Thus the Lagrangian is<br />

= 1 2 ( +2)( ˙2 + ˙ 2 + ˙ 2 )+( ˙ 2 + 2 ˙2 + 2 ˙ 2 s<strong>in</strong> 2 )+ 1 24 2 (˙ 2 + ˙ 2 s<strong>in</strong> 2 ) − ( − 0 ) 2<br />

Us<strong>in</strong>g Lagrange’s equations Λ =0for the generalized coord<strong>in</strong>ates gives.<br />

O<br />

z<br />

1<br />

r<br />

y<br />

1


6.9. APPLICATIONS INVOLVING NON-HOLONOMIC CONSTRAINTS 157<br />

( +2) ˙ = constant (Λ =0)<br />

( +2) ˙ = constant (Λ =0)<br />

( +2) ˙ = constant (Λ =0)<br />

µ2 2 + 1 <br />

12 2 ˙ s<strong>in</strong> 2 = constant (Λ =0)<br />

¨ − ˙ 2 − ˙ 2 s<strong>in</strong> 2 + ( − 0) = 0 (Λ =0)<br />

<br />

<br />

µ 2 + 2 ¨ +2 ˙<br />

24<br />

˙ −<br />

µ 2 + 2 ˙ 2 s<strong>in</strong> cos = 0 (Λ =0)<br />

24<br />

The first three equations show that the three components of the l<strong>in</strong>ear momentum of the center of mass<br />

are constants of motion. The fourth equation shows that the component of the angular momentum about<br />

the 0 axis is a constant of motion. S<strong>in</strong>ce the 1 axis has been arbitrarily chosen then the total angular<br />

momentum must be conserved. The fifth and sixth equations give the radial and angular equations of motion<br />

of the oscillat<strong>in</strong>g masses .<br />

6.9 Applications <strong>in</strong>volv<strong>in</strong>g non-holonomic constra<strong>in</strong>ts<br />

In general, non-holonomic constra<strong>in</strong>ts can be handled by use of generalized forces <br />

<strong>in</strong> the Lagrange-<br />

Euler equations 660. The follow<strong>in</strong>g examples, 616 − 619 <strong>in</strong>volve one-sided constra<strong>in</strong>ts which exhibit<br />

holonomic behavior for restricted ranges of the constra<strong>in</strong>t surface <strong>in</strong> coord<strong>in</strong>ate space, and this range is case<br />

specific. When the forces of constra<strong>in</strong>t press the object aga<strong>in</strong>st the constra<strong>in</strong>t surface, then the system is<br />

holonomic, but the holonomic range of coord<strong>in</strong>ate space is limited to situations where the constra<strong>in</strong>t forces<br />

are positive. When the constra<strong>in</strong>t force is negative, the object flies free from the constra<strong>in</strong>t surface. In<br />

addition, when the frictional force where is the static coefficient of friction, then the<br />

object slides negat<strong>in</strong>g any roll<strong>in</strong>g constra<strong>in</strong>t that assumes static friction.<br />

6.16 Example: Mass slid<strong>in</strong>g on a frictionless spherical shell<br />

Consider a mass starts from rest at the top of a frictionless<br />

fixed spherical shell of radius . The questions are what is the<br />

force of constra<strong>in</strong>t and determ<strong>in</strong>e the angle at which the mass<br />

leaves the surface of the spherical shell. The coord<strong>in</strong>ates shown<br />

are the obvious generalized coord<strong>in</strong>ates to use. The constra<strong>in</strong>t will<br />

not apply if the force of constra<strong>in</strong>t does not hold the mass aga<strong>in</strong>st<br />

the surface of the spherical shell, that is, it is only holonomic <strong>in</strong> a<br />

restricted doma<strong>in</strong>.<br />

The Lagrangian is<br />

= 1 2 ³<br />

˙ 2 + 2 ˙2´ − cos <br />

This Lagrangian is applicable irrespective of whether the constra<strong>in</strong>t<br />

is obeyed, where the constra<strong>in</strong>t is given by<br />

( ) = − =0<br />

Mass slid<strong>in</strong>g on frictionless cyl<strong>in</strong>der<br />

of radius .<br />

For the restricted doma<strong>in</strong> where this system is holonomic, it can be solved us<strong>in</strong>g generalized coord<strong>in</strong>ates,<br />

generalized forces, Lagrange multipliers, or Newtonian mechanics as illustrated below.<br />

M<strong>in</strong>imal generalized coord<strong>in</strong>ates:<br />

The m<strong>in</strong>imal number of generalized coord<strong>in</strong>ates reduces the system to one coord<strong>in</strong>ate , which does not<br />

determ<strong>in</strong>e the constra<strong>in</strong>t force that is needed to know if the constra<strong>in</strong>t applies. Thus this approach is not<br />

useful for solv<strong>in</strong>g this partially-holonomic system.


158 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

Generalized forces:<br />

The radial constra<strong>in</strong>t has a correspond<strong>in</strong>g generalized force . The Lagrange equation Λ = <br />

¨ + cos − ˙ 2 = <br />

gives<br />

(a)<br />

TheLagrangeequationΛ = =0s<strong>in</strong>ce there is no tangential force for this frictionless system. Therefore<br />

2¨ − s<strong>in</strong> +2 ˙ ˙ =0<br />

When constra<strong>in</strong>ed to follow the surface of the spherical shell, the system is holonomic, i.e.<br />

˙ =¨ =0. Thus the above two equations reduce to<br />

(b)<br />

= and<br />

That is<br />

Integrate to get ˙ us<strong>in</strong>g the fact that<br />

then<br />

Z<br />

cos − ˙ 2 = (c)<br />

2¨ − s<strong>in</strong> = 0<br />

¨ =<br />

<br />

s<strong>in</strong> <br />

¨ = ˙ ˙<br />

= ˙<br />

<br />

Z<br />

¨ = ˙˙ = Z<br />

s<strong>in</strong> <br />

<br />

Therefore<br />

˙ 2 = 2 <br />

(1 − cos )<br />

(d)<br />

assum<strong>in</strong>g that ˙ =0at =0 Substitut<strong>in</strong>g equation () <strong>in</strong>to equation () gives the constra<strong>in</strong>t force, which<br />

is normal to the surface, to be<br />

= = (3 cos − 2)<br />

Note that = =0when cos = 2 3 , that is =482 <br />

Lagrange multipliers:<br />

For the holonomic regime, which obeys the constra<strong>in</strong>t, ( ) = − =0 the Lagrange equation for <br />

is Λ = <br />

<br />

S<strong>in</strong>ce<br />

=1 then ¨ + cos − ˙ 2 = (a)<br />

The Lagrange equation for gives ∆ = <br />

<br />

<br />

=0s<strong>in</strong>ce<br />

<br />

=0 Thus<br />

2¨ − s<strong>in</strong> +2 ˙ ˙ =0<br />

(b)<br />

As above, when constra<strong>in</strong>ed to follow the surface of the spherical shell, the system is holonomic = <br />

and ˙ =¨ =0 Thus the above two equations reduce to<br />

cos − ˙ 2 = (c)<br />

2¨ − s<strong>in</strong> = 0 (d)<br />

That is, the answers are identical to that obta<strong>in</strong>ed us<strong>in</strong>g generalized forces, namely;<br />

˙ 2 = 2 <br />

(1 − cos )<br />

(d)<br />

assum<strong>in</strong>g that ˙ =0at =0<br />

The force of constra<strong>in</strong>t applied by the surface is<br />

= <br />

=


6.9. APPLICATIONS INVOLVING NON-HOLONOMIC CONSTRAINTS 159<br />

Substitut<strong>in</strong>g equation () <strong>in</strong>to equation () gives<br />

= = (3 cos − 2)<br />

Note that =0when cos = 2 3 , that is =482 <br />

Both of the above methods give identical results and give that the force of constra<strong>in</strong>t is negative when<br />

482 Assum<strong>in</strong>g that the surface cannot hold the mass aga<strong>in</strong>st the surface, then the mass will fly off the<br />

spherical shell when 482 and the system reduces to an unconstra<strong>in</strong>ed object fall<strong>in</strong>g freely <strong>in</strong> a uniform<br />

gravitational field, which is holonomic, that is = =0 Then the equations of motion () and () reduce<br />

to<br />

Energy conservation:<br />

This problem can be solved us<strong>in</strong>g energy conservation<br />

¨ + cos − ˙ 2 = 0 (e)<br />

2¨ − s<strong>in</strong> +2 ˙ ˙ = 0 (f)<br />

1<br />

2 2 = [1 − cos ]<br />

Thus the centripetal acceleration<br />

2<br />

=2[1 − cos ]<br />

<br />

The normal force to the surface will cancel when the centripetal acceleration equals the gravitational acceleration,<br />

that is, when<br />

2<br />

=2[1 − cos ] = cos <br />

<br />

This occurs when cos = 2 3<br />

. This is an unusual case where the Newtonian approach is the simplest.<br />

6.17 Example: Roll<strong>in</strong>g solid sphere on a spherical shell<br />

This is a similar problem to the prior one with the added<br />

complication of roll<strong>in</strong>g which is assumed to move <strong>in</strong> a vertical<br />

plane mak<strong>in</strong>g it holonomic. Here we would like to determ<strong>in</strong>e<br />

the forces of constra<strong>in</strong>t to see when the solid sphere flies off the<br />

spherical shell and when the friction is <strong>in</strong>sufficient to stop the<br />

roll<strong>in</strong>gspherefromslipp<strong>in</strong>g.<br />

The best generalized coord<strong>in</strong>ates are the distance of the center<br />

of the sphere from the center of the spherical shell, and <br />

It is important to note that is measured with respect to the<br />

vertical, not the time-dependent vector r. That is, the direction<br />

of the radius is whichistimedependentandthusisnota<br />

useful reference to use to def<strong>in</strong>e the angle . Let us assume<br />

that the sphere is uniform with a moment of <strong>in</strong>ertia of =<br />

2<br />

5 2 If the tangential frictional force is less than the limit<strong>in</strong>g<br />

value ,with 0 then the sphere will roll without<br />

slipp<strong>in</strong>g on the surface of the cyl<strong>in</strong>der and both constra<strong>in</strong>ts apply.<br />

Disk of mass , radius roll<strong>in</strong>g on a<br />

cyl<strong>in</strong>drical surface of radius .<br />

Under these conditions the system is holonomic and the solution is solved us<strong>in</strong>g Lagrange multipliers and the<br />

equations of constra<strong>in</strong>t are the follow<strong>in</strong>g:<br />

1) The center of the sphere follows the surface of the cyl<strong>in</strong>der<br />

2) The sphere rolls without slipp<strong>in</strong>g<br />

1 = − − =0<br />

2 = ( − ) − =0


160 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

³<br />

The k<strong>in</strong>etic energy is = 1 2 ˙ 2 + 2 ˙ 2´ + 1 2 ˙ 2 and the potential energy is = cos Thus the<br />

Lagrangian is<br />

= 1 ³<br />

2 ˙ 2 + ˙2´ 2 + 1 2 ˙ 2 − cos <br />

Consider the solution us<strong>in</strong>g Lagrange multipliers for the holonomic regime where both constra<strong>in</strong>ts are<br />

satisfied and lead to the follow<strong>in</strong>g differential constra<strong>in</strong>t relations<br />

1<br />

<br />

= 1<br />

2<br />

<br />

= 0<br />

The Lagrange operator equation Λ gives,<br />

that is<br />

Λ gives<br />

1<br />

=0 1<br />

=0<br />

2<br />

= 2<br />

= − ( + )<br />

<br />

<br />

˙ − <br />

= 1<br />

1<br />

+ 2<br />

2<br />

<br />

¨ + cos − ˙ 2 = 1<br />

2¨ +2 ˙ ˙ − s<strong>in</strong> = −2 ( + )<br />

Λ gives<br />

¨ = 2<br />

S<strong>in</strong>ce the center of the sphere roll<strong>in</strong>g on the spherical shell must have<br />

(a)<br />

(b)<br />

(c)<br />

= + <br />

then<br />

˙ = ¨ =0<br />

¨ =<br />

¨<br />

<br />

Substitut<strong>in</strong>g this <strong>in</strong>to () gives<br />

<br />

¨ 2<br />

=<br />

2<br />

Insert this <strong>in</strong>to equation () gives<br />

s<strong>in</strong> <br />

2 = ¡ ¢<br />

+<br />

2 2<br />

<br />

The moment of <strong>in</strong>ertia about the axis of a solid sphere is = 2 5 2 Then<br />

2 s<strong>in</strong> <br />

2 =<br />

7<br />

But also<br />

˙<br />

¨ = ˙<br />

= 2<br />

2 = 5<br />

2 5 s<strong>in</strong> <br />

2 =<br />

7<br />

Integrat<strong>in</strong>g gives<br />

Z<br />

˙˙ = 5 Z<br />

s<strong>in</strong> <br />

7<br />

That is<br />

˙ 2 = 10 (1 − cos )<br />

7<br />

assum<strong>in</strong>g that ˙ =0at =0 Insert<strong>in</strong>g this <strong>in</strong>to equation () gives<br />

− 10<br />

7 [1 − cos ]+ cos = 1


6.9. APPLICATIONS INVOLVING NON-HOLONOMIC CONSTRAINTS 161<br />

That is<br />

1 = [17 cos − 10]<br />

7<br />

Note that this equals zero when<br />

cos = 10<br />

17<br />

For larger angles 1 is negative imply<strong>in</strong>g that the solid sphere will fly off the surface of the spherical shell.<br />

Thespherewillleavethesurfaceofthecyl<strong>in</strong>derwhencos = 10<br />

17 that is, =5397 This is a significantly<br />

larger angle than obta<strong>in</strong>ed for the similar problem where the mass is slid<strong>in</strong>g on a frictionless cyl<strong>in</strong>der because<br />

the energy stored <strong>in</strong> rotation implies that the l<strong>in</strong>ear velocity of the mass is lower at a given angle for the<br />

case of a roll<strong>in</strong>g sphere.<br />

The above discussion has omitted an important fact that, if ∞ the frictional force becomes<br />

<strong>in</strong>sufficient to ma<strong>in</strong>ta<strong>in</strong> the roll<strong>in</strong>g constra<strong>in</strong>t before = 5397 that is, the frictional force will exceed<br />

the slid<strong>in</strong>g limit . To determ<strong>in</strong>e when the roll<strong>in</strong>g constra<strong>in</strong>t fails it is necessary to determ<strong>in</strong>e the<br />

frictional torque<br />

= − 2 <br />

Thus<br />

= − 2<br />

It is <strong>in</strong> the negative direction because of the direction chosen for The required coefficient of friction is<br />

given by the ratio of the frictional force to the normal force, that is<br />

= 2 2s<strong>in</strong><br />

=<br />

1 [17 cos − 10]<br />

For =1the disk starts to slip when =4754 0 Note that the sphere starts slipp<strong>in</strong>g before it flies off<br />

the cyl<strong>in</strong>der s<strong>in</strong>ce a normal force is required to support a frictional force and the difference depends on the<br />

coefficient of friction. The no-slipp<strong>in</strong>g constra<strong>in</strong>t is not satisfied once the sphere starts slipp<strong>in</strong>g and the<br />

frictional force should equal 1 Thusfortheanglesbeyond4754 theproblemneedstobesolvedwith<br />

the roll<strong>in</strong>g constra<strong>in</strong>t changed to a slid<strong>in</strong>g non-conservative frictional force. This is best handled by <strong>in</strong>clud<strong>in</strong>g<br />

the frictional force and normal forces as generalized forces. Fortunately this will be a small correction. The<br />

friction will slightly change the exact angle at which the normal force becomes zero and the system transitions<br />

to free motion of the sphere <strong>in</strong> a gravitational field.<br />

6.18 Example: Solid sphere roll<strong>in</strong>g plus slipp<strong>in</strong>g on a spherical shell<br />

Consider the above case when the frictional force is <strong>in</strong>sufficient to constra<strong>in</strong> the motion to roll<strong>in</strong>g. Now<br />

the frictional force is given by<br />

= <br />

when is positive.<br />

This can be solved us<strong>in</strong>g generalized forces with the previous Lagrangian. Then<br />

<br />

˙ − <br />

= = <br />

which gives<br />

¨ + cos − ˙ 2 = <br />

Similarly Λ = = − ( + ) gives<br />

Similarly Λ = = gives<br />

2¨ +2 ˙ ˙ − s<strong>in</strong> = − ( + )<br />

¨ = <br />

These can be solved by substitut<strong>in</strong>g the relation = . The sphere flies off the spherical shell<br />

when ≤ 0 lead<strong>in</strong>g to free motion discussed <strong>in</strong> example 62. The problem of a solid uniform sphere roll<strong>in</strong>g<br />

<strong>in</strong>side a hollow sphere can be solved the same way.


162 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

6.19 Example: Small body held by friction on the periphery of a roll<strong>in</strong>g wheel<br />

Assume that a small body of mass is balanced<br />

on a roll<strong>in</strong>g wheel of mass and radius<br />

as shown <strong>in</strong> the figure. The wheel rolls <strong>in</strong><br />

a vertical plane without slipp<strong>in</strong>g on a horizontal<br />

surface. This example illustrates that it is possible<br />

to use simultaneously a mixture of holonomic<br />

constra<strong>in</strong>ts, partially-holonomic constra<strong>in</strong>ts, and<br />

generalized forces. 3<br />

Assume that at =0the wheel touches the<br />

floor at = = 0 with the mass perched at<br />

the top of the wheel at =0. Let the frictional<br />

force act<strong>in</strong>g on the mass be and the reaction<br />

force of the periphery of the wheel on the mass<br />

be . Let ˙ be the angular velocity of the wheel,<br />

and ˙ the horizontal velocity of the center of the<br />

wheel. The polar coord<strong>in</strong>ates of the mass <br />

are taken with measured from the center of the<br />

wheel with measured with respect to the vertical.<br />

Thus the cartesian coord<strong>in</strong>ates of the small mass<br />

are ( + s<strong>in</strong> + cos ) with respect to the<br />

orig<strong>in</strong> at = =0.<br />

The k<strong>in</strong>etic energy is given by<br />

y<br />

F<br />

M<br />

x<br />

O<br />

x<br />

Small body of mass held by friction on the periphery<br />

of a roll<strong>in</strong>g wheel of mass and radius .<br />

= 1 2 ˙2 + 1 2 ˙2 + 1 ∙ ³<br />

2 ˙ + ˙<br />

´2 ³<br />

´2¸<br />

cos + ˙ s<strong>in</strong> + ˙ cos − ˙ s<strong>in</strong> <br />

m<br />

N<br />

The gravitational force can be absorbed <strong>in</strong>to the scalar potential term of the Lagrangian and <strong>in</strong>cludes only<br />

the potential energy of the mass s<strong>in</strong>ce the potential energy of the roll<strong>in</strong>g wheel is constant.<br />

Thus the Lagrangian is<br />

=+ ( + cos )<br />

= 1 2 ( + ) ˙2 + 1 2 ˙2 + 1 2 h 2 ˙2 +2 ˙˙ cos +2˙ ˙ s<strong>in</strong> + ˙<br />

2 i − ( + cos )<br />

The equations of constra<strong>in</strong>ts are:<br />

1) The wheel rolls without slipp<strong>in</strong>g on the ground plane lead<strong>in</strong>g to a holonomic constra<strong>in</strong>t:<br />

1 = − = ˙ − ˙ =0<br />

2) The mass is touch<strong>in</strong>g the periphery of the wheel, that is, the normal force 0 This is a one-sided<br />

restricted holonomic constra<strong>in</strong>t.<br />

2 = − =0<br />

3) The mass does not slip on the wheel if the frictional force . When this restricted<br />

holonomic constra<strong>in</strong>t is satisfied, then<br />

3 = ˙ − ˙ =0<br />

The roll<strong>in</strong>g constra<strong>in</strong>t is holonomic, and can be accounted for us<strong>in</strong>g one Lagrange multiplier plus the<br />

differential constra<strong>in</strong>t equations<br />

3 This problem is solved <strong>in</strong> detail <strong>in</strong> example 319 of " <strong>Classical</strong> <strong>Mechanics</strong> and Relativity". by Muller-Kirsten [06]


6.9. APPLICATIONS INVOLVING NON-HOLONOMIC CONSTRAINTS 163<br />

1<br />

<br />

= 1<br />

1<br />

<br />

= 0<br />

1<br />

= <br />

1<br />

<br />

= 0<br />

The other two constra<strong>in</strong>ts are non-holonomic, and thus these constra<strong>in</strong>t forces are expressed <strong>in</strong> terms of two<br />

generalized forces and that are related to the tangential force and radial reaction force . For<br />

simplicity, assume that the wheel is a th<strong>in</strong>-walled cyl<strong>in</strong>der with a moment of <strong>in</strong>ertia of<br />

= 2<br />

The Euler-Lagrange equations for the four coord<strong>in</strong>ates are<br />

− ³<br />

( + ) ˙ + <br />

<br />

´<br />

cos + ˙ s<strong>in</strong> + + = 0 (Λ )<br />

˙˙ s<strong>in</strong> + ˙ ˙ cos − s<strong>in</strong> − ³ 2 ˙ + ˙ cos ´<br />

+ = 0 (Λ )<br />

<br />

− ¡<br />

2 ˙ ¢ − <br />

<br />

= 0 (Λ )<br />

− cos − ( ˙ s<strong>in</strong> + ˙)+ = 0 (Λ )<br />

The generalized forces can be related to and us<strong>in</strong>g the def<strong>in</strong>ition<br />

= F()· r<br />

<br />

where () is the vectorial sum of the forces act<strong>in</strong>g at The components of vector =( + s<strong>in</strong> + cos )<br />

and ,and are <strong>in</strong> the directions def<strong>in</strong>ed <strong>in</strong> the figure which leads to the generalized forces<br />

= − cos + s<strong>in</strong> <br />

= (− cos + s<strong>in</strong> )(− cos ) − ( s<strong>in</strong> + cos ) s<strong>in</strong> = −<br />

= <br />

Solv<strong>in</strong>g the above 7 equations gives that<br />

¨ s<strong>in</strong> + ˙ 2 − cos + =0<br />

This last equation can be derived by Newtonian mechanics from consideration of the forces act<strong>in</strong>g.<br />

The above equations of motion can be used to calculate the motion for the follow<strong>in</strong>g conditions.<br />

a) Mass not slipp<strong>in</strong>g:<br />

This occurs if = ≤ whichalsoimpliesthat0 That is a situation where the system is<br />

holonomic with = ˙ = ˙ ˙ = ˙ which can be solved us<strong>in</strong>g the generalized coord<strong>in</strong>ate approach with<br />

only one <strong>in</strong>dependent coord<strong>in</strong>ate which can be taken to be .<br />

b) Mass slipp<strong>in</strong>g:<br />

Here the no-slip constra<strong>in</strong>t is violated and thus one has to explicitly <strong>in</strong>clude the generalized forces <br />

and assume that slid<strong>in</strong>g friction is given by = <br />

c) Reaction force is negative:<br />

Here the mass is not subject to any constra<strong>in</strong>ts and it is <strong>in</strong> free fall.<br />

The above example illustrates the flexibility provided by Lagrangian mechanics that allows simultaneous<br />

use of Lagrange multipliers, generalized forces, and scalar potential to handle comb<strong>in</strong>ations of several<br />

holonomic and nonholonomic constra<strong>in</strong>ts for a complicated problem.


164 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

6.10 Velocity-dependent Lorentz force<br />

The Lorentz force <strong>in</strong> electromagnetism is unusual <strong>in</strong> that it is a velocity-dependent force, as well as be<strong>in</strong>g a<br />

conservative force that can be treated us<strong>in</strong>g the concept of potential. That is, the Lorentz force is<br />

F = (E + v × B) (6.61)<br />

It is <strong>in</strong>terest<strong>in</strong>g to use Maxwell’s equations and Lagrangian mechanics to show that the Lorentz force can be<br />

represented by a conservative potential <strong>in</strong> Lagrangian mechanics.<br />

Maxwell’s equations can be written as<br />

∇ · E = 0<br />

(6.62)<br />

∇ × E+ B<br />

<br />

= 0<br />

∇ · B = 0<br />

E<br />

∇ × B− 0 0<br />

<br />

= J<br />

S<strong>in</strong>ce ∇ · B =0 then it follows from Appendix that B can be represented by the curl of a vector<br />

potential, A that is<br />

B = ∇ × A (6.63)<br />

Substitut<strong>in</strong>g this <strong>in</strong>to ∇ × E+ B<br />

<br />

=0gives that<br />

∇ × E+ ∇ × A<br />

µ<br />

<br />

∇× E + A <br />

<br />

= 0 (6.64)<br />

S<strong>in</strong>ce this curl is zero it can be represented by the gradient of a scalar potential <br />

E + A = −∇ (6.65)<br />

<br />

The follow<strong>in</strong>g shows that this relation corresponds to tak<strong>in</strong>g the gradient of a potential for the charge <br />

where the potential is given by the relation<br />

= 0<br />

= (Φ − A · v) (6.66)<br />

where Φ is the scalar electrostatic potential. This scalar potential can be employed <strong>in</strong> the Lagrange<br />

equations us<strong>in</strong>g the Lagrangian<br />

= 1 v · v − (Φ − A · v) (6.67)<br />

2<br />

The Lorentz force can be derived from this Lagrangian by consider<strong>in</strong>g the Lagrange equation for the cartesian<br />

coord<strong>in</strong>ate <br />

Us<strong>in</strong>g the above Lagrangian (667) gives<br />

¨ + <br />

<br />

˙ − =0 (6.68)<br />

<br />

∙ <br />

<br />

+ Φ<br />

− A<br />

· v¸<br />

=0 (6.69)<br />

But<br />

and<br />

<br />

<br />

= <br />

<br />

+ <br />

˙ + <br />

˙ + ˙ (6.70)<br />

<br />

A<br />

· v = <br />

˙ + <br />

˙ + ˙ (6.71)


6.11. TIME-DEPENDENT FORCES 165<br />

Insert<strong>in</strong>g equations 670 and 671 <strong>in</strong>to 669 gives<br />

∙µ<br />

= ¨ = − Φ<br />

− µ<br />

<br />

+<br />

− <br />

<br />

<br />

µ <br />

˙ −<br />

<br />

− ¸<br />

<br />

˙ = [E + v × B]<br />

<br />

<br />

(6.72)<br />

Correspond<strong>in</strong>g expressions can be obta<strong>in</strong>ed for and . Thus the total force is the well-known Lorentz<br />

force<br />

F = (E + v × B) (6.73)<br />

This has demonstrated that the electromagnetic scalar potential<br />

= (Φ − A · v) (6.74)<br />

satisfies Maxwell’s equations, gives the Lorentz force, and it can be absorbed <strong>in</strong>to the Lagrangian. Note that<br />

the velocity-dependent Lorentz force is conservative s<strong>in</strong>ce E is conservative, and because (v × B × v)=0<br />

therefore the magnetic force does no work s<strong>in</strong>ce it is perpendicular to the trajectory. The velocity-dependent<br />

conservative Lorentz force is an important and ubiquitous force that features prom<strong>in</strong>ently <strong>in</strong> many branches<br />

of science. It will be discussed further for the case of relativistic motion <strong>in</strong> example 176.<br />

6.11 Time-dependent forces<br />

All examples discussed <strong>in</strong> this chapter have assumed Lagrangians that are time <strong>in</strong>dependent. Mathematical<br />

systems where the ord<strong>in</strong>ary differential equations do not depend explicitly on the <strong>in</strong>dependent variable, which<br />

<strong>in</strong> this case is time , are called autonomous systems. Systems hav<strong>in</strong>g differential equations govern<strong>in</strong>g the<br />

dynamical behavior that have time-dependent coefficients are called non-autonomous systems.<br />

In pr<strong>in</strong>ciple it is trivial to <strong>in</strong>corporate time-dependent behavior <strong>in</strong>to the equations of motion by <strong>in</strong>troduc<strong>in</strong>g<br />

either a time dependent generalized force ( ), or allow<strong>in</strong>g the Lagrangian to be time dependent.<br />

For example, <strong>in</strong> the rocket problem the mass is time dependent. In some cases the time dependent forces<br />

can be represented by a time-dependent potential energy rather than us<strong>in</strong>g a generalized force. Solutions<br />

for non-autonomous systems can be considerably more difficult to obta<strong>in</strong>, and can <strong>in</strong>volve regions where the<br />

motion is stable and other regions where the motion is unstable or chaotic similar to the behavior discussed<br />

<strong>in</strong> chapter 4. The follow<strong>in</strong>g case of a simple pendulum, whose support is undergo<strong>in</strong>g vertical oscillatory<br />

motion, illustrates the complexities that can occur for systems <strong>in</strong>volv<strong>in</strong>g time-dependent forces.<br />

6.20 Example: Plane pendulum hang<strong>in</strong>g from a vertically-oscillat<strong>in</strong>g support<br />

Consider a plane pendulum hav<strong>in</strong>g a mass fastened to a massless rigid rod of length that is at an<br />

angle () to the vertical gravitational field . The pendulum is attached to a support that is subject to a<br />

vertical oscillatory force such that the vertical position of the support is<br />

= cos <br />

The k<strong>in</strong>etic energy is<br />

= 1 ∙ ³ ´2<br />

2 ˙ cos +(˙ + ˙ s<strong>in</strong> )<br />

2¸<br />

= 1 h i<br />

2 2 ˙2 +2˙ ˙ s<strong>in</strong> + ˙<br />

2<br />

and the potential energy is<br />

Thus the Lagrangian is<br />

= [(1 − cos )+]<br />

= 1 h i 2 2 ˙2 +2˙ ˙ s<strong>in</strong> + ˙<br />

2<br />

− [(1 − cos )+]<br />

The Euler-Lagrange equations lead to equations of motion for and <br />

2¨ + ¨ s<strong>in</strong> + s<strong>in</strong> = 0<br />

¨ s<strong>in</strong> + ˙ 2 cos + ¨ + =


166 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

Assume the small-angle approximation where → 0 then these two equations reduce to<br />

µ <br />

¨ +<br />

<br />

+ ¨ = 0<br />

¨ + = <br />

Substitute ¨ = − 2 cos <strong>in</strong>to these equations gives<br />

µ <br />

<br />

¨ +<br />

− 2<br />

cos = 0<br />

¡ − 2 cos ¢ = <br />

These correspond to stable harmonic oscillations about ≈ 0 if the bracket term is positive, and to<br />

unstable motion if the bracket is negative. Thus, for small amplitude oscillation about ≈ 0 the motion of<br />

the system can be unstable whenever the bracket is negative, that is, when the acceleration 2 cos <br />

and resonance behavior can occur coupl<strong>in</strong>g the pendulum period and the forc<strong>in</strong>g frequency .<br />

This discussion also applies to the <strong>in</strong>verted pendulum with a surpris<strong>in</strong>g result. It is well known that the<br />

pendulum is unstable near = . However, if the support is oscillat<strong>in</strong>g, then for ≈ the equations of<br />

motion become<br />

µ <br />

<br />

¨ −<br />

− 2<br />

cos = 0<br />

¡ − 2 cos ¢ = <br />

The <strong>in</strong>verted pendulum has stable oscillations about ≈ if the bracket is negative, that is, if 2 cos <br />

This illustrates that nonautonomous dynamical systems can <strong>in</strong>volve either stable or unstable motion.<br />

6.12 Impulsive forces<br />

Collid<strong>in</strong>g bodies often <strong>in</strong>volve large impulsive forces that act for a short time. As discussed <strong>in</strong> chapter 2128<br />

the treatment of impulsive forces or torques is greatly simplified if they act for a sufficiently short time that<br />

the displacement dur<strong>in</strong>g the impact can be ignored, even though the <strong>in</strong>stantaneous change <strong>in</strong> velocities may<br />

be large. The simplicity is achieved by tak<strong>in</strong>g the time <strong>in</strong>tegral of the Euler-Lagrange equations over the<br />

duration of the impulse and assum<strong>in</strong>g → 0.<br />

The impact of the impulse on a system can be handled two ways. The first approach is to use the<br />

Euler-Lagrange equation dur<strong>in</strong>g the impulse to determ<strong>in</strong>e the equations of motion<br />

<br />

<br />

µ <br />

˙ <br />

<br />

− <br />

<br />

= <br />

(6.75)<br />

where the impulsive force is <strong>in</strong>troduced us<strong>in</strong>g the generalized force <br />

. Know<strong>in</strong>g the <strong>in</strong>itial conditions at<br />

time the conditions at the time + are given by <strong>in</strong>tegration of equation 675 over the duration of the<br />

impulse which gives<br />

Z +<br />

<br />

µ Z<br />

<br />

+<br />

−<br />

˙ <br />

Z<br />

<br />

+<br />

= <br />

(6.76)<br />

<br />

This <strong>in</strong>tegration determ<strong>in</strong>es the conditions at time + which then are used as the <strong>in</strong>itial conditions for the<br />

motion when the impulsive force <br />

is zero.<br />

The second approach is to realize that equation 676 can be rewritten <strong>in</strong> the form<br />

Z +<br />

lim<br />

→0<br />

<br />

µ <br />

<br />

<br />

= lim<br />

˙ →0 ˙ <br />

¯¯¯¯<br />

+<br />

<br />

Z +<br />

µµ <br />

<br />

= ∆ = lim<br />

+ <br />

(6.77)<br />

→0<br />

<br />

Note that <strong>in</strong> the limit that → 0 then the <strong>in</strong>tegral of the generalized momentum = <br />

˙ <br />

simplifies to give<br />

³ ´<br />

<br />

the change <strong>in</strong> generalized momentum ∆ . In addition, assum<strong>in</strong>g that the non-impulsive forces<br />

<br />

are


6.12. IMPULSIVE FORCES 167<br />

f<strong>in</strong>ite and <strong>in</strong>dependent of the <strong>in</strong>stantaneous impulsive force dur<strong>in</strong>g the <strong>in</strong>f<strong>in</strong>itessimal duration , then the<br />

contribution of the non-impulsive forces R ³ ´<br />

+ <br />

<br />

dur<strong>in</strong>g the impulse can be neglected relative to the<br />

large impulsive force term; lim →0<br />

R +<br />

<br />

<br />

. Thus it can be assumed that<br />

Z +<br />

∆ = lim <br />

= ˜ (6.78)<br />

→0<br />

<br />

where ˜ is the generalized impulse associated with coord<strong>in</strong>ate =1 2 3 . This generalized impulse<br />

can be derived from the time <strong>in</strong>tegral of the impulsive forces P given by equation 2135 us<strong>in</strong>g the time<br />

<strong>in</strong>tegral of equation 677, thatis<br />

Z +<br />

∆ = ˜ =lim<br />

→0<br />

<br />

Z +<br />

<br />

≡ lim<br />

→0<br />

<br />

X<br />

<br />

P · r <br />

<br />

= X <br />

˜P · r <br />

<br />

(6.79)<br />

Note that the generalized impulse ˜ can be a translational impulse ˜P with correspond<strong>in</strong>g translational<br />

variable or an angular impulsive torque ˜τ with correspond<strong>in</strong>g angular variable .<br />

Impulsive force problems usually are solved <strong>in</strong> two stages. Either equations 676 or 679 are used to<br />

determ<strong>in</strong>e the conditions of the system immediately follow<strong>in</strong>g the impulse. If → 0 then impulse changes<br />

the generalized velocities ˙ but not the generalized coord<strong>in</strong>ates . The subsequent motion then is determ<strong>in</strong>ed<br />

us<strong>in</strong>g the Lagrangian equations of motion with the impulsive generalized force be<strong>in</strong>g zero, and assum<strong>in</strong>g that<br />

the <strong>in</strong>itial condition corresponds to the result of the impulse calculation.<br />

6.21 Example: Series-coupled double pendulum subject to impulsive force<br />

Consider a series-coupled double pendulum compris<strong>in</strong>g<br />

two masses 1 and 2 connected by rigid massless rods of<br />

lengths 1 and 2 asshown<strong>in</strong>thefigure. Initially the two<br />

pendula are at rest and hang<strong>in</strong>g vertically when a horizontal<br />

impulse ˜ strikes the system at a distance below the upper<br />

fulcrum where 1 1 + 2 . For this system the<br />

k<strong>in</strong>etic energy of the masses 1 and 2 are<br />

1 = 1 2 1 2 1 ˙ 2 1<br />

2 = 1 2 2[ 2 1 ˙ 2 1 +2 1 2 ˙1 ˙2 cos( 1 − 2 )+ 2 2 ˙ 2 2]<br />

Note the velocity of 2 is the vector sum of the two velocities<br />

shown, separated by the angle 2 − 1 . Thus the total k<strong>in</strong>etic<br />

energy is<br />

Two series-coupled plane pendula.<br />

To first order <strong>in</strong> cos( 1 − 2 )<br />

= 1 2 ( 1 + 2 ) 2 1 ˙ 2 1 + 2 1 2 ˙1 ˙2 cos( 1 − 2 )+ 1 2 2 2 2 ˙ 2 2<br />

= 1 2 ( 1 + 2 ) 2 ˙ 2 1 1 + 2 1 2 ˙1 ˙2 + 1 2 2 2 ˙ 2 2 2<br />

The total potential energy is<br />

= 1 1 (1 − cos 1 )+ 2 [ 1 (1 − cos 1 )+ 2 (1 − cos 2 )<br />

= ( 1 + 2 ) 1 (1 − cos 1 )+ 2 2 (1 − cos 2 )<br />

Thus, assum<strong>in</strong>g the small-angle approximation, the Lagrangian becomes<br />

<br />

= 1 2 ( 1 + 2 ) 2 1 ˙ 2 1 + 2 1 2 ˙1 ˙2 + 1 2 2 2 2 ˙ 2 2 −<br />

µ 1<br />

2 ( 1 + 2 ) 1 2 1 + 1 2 2 2 2 2


168 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

Use equation 679 to transform to the generalized coord<strong>in</strong>ates 1 and 2 with the correspond<strong>in</strong>g generalized<br />

impulsive torques<br />

˜ 1 = ˜ 1<br />

˜ 2 = ˜ ( − 1 )<br />

S<strong>in</strong>ce the system starts at rest where 1 = 2 =0, then us<strong>in</strong>g equation 677 gives the change <strong>in</strong> angular<br />

momentum immediately follow<strong>in</strong>g the impulse to be<br />

1 2 ˙<br />

³<br />

1 1 + 2 1 1 ˙1 + 2 ˙2´<br />

2 2<br />

³<br />

1 ˙1 + 2 ˙2´<br />

= ˜ 1<br />

= ˜ ( − 1 )<br />

These two equations determ<strong>in</strong>e ˙ 1 and ˙ 2 immediately after the impulse; these can be used with 1 = 2 =0<br />

as <strong>in</strong>itial conditions for solv<strong>in</strong>g the subsequent force-free motion when the generalized impulsive force is zero.<br />

As described <strong>in</strong> example 145 the subsequent motion of this series coupled pendulum will be a superposition<br />

of the two normal modes with amplitudes determ<strong>in</strong>ed by the result of the impulse calculation.<br />

6.13 TheLagrangianversustheNewtonianapproachtoclassical<br />

mechanics<br />

It is useful to contrast the differences, and relative advantages, of the Newtonian and Lagrangian formulations<br />

of classical mechanics. The Newtonian force-momentum formulation is vectorial <strong>in</strong> nature, it has cause and<br />

effect embedded <strong>in</strong> it. The Lagrangian approach is cast <strong>in</strong> terms of k<strong>in</strong>etic and potential energies which <strong>in</strong>volve<br />

only scalar functions and the equations of motion come from a s<strong>in</strong>gle scalar function, i.e. Lagrangian. The<br />

directional properties of the equations of motion come from the requirement that the trajectory is specified<br />

by the pr<strong>in</strong>ciple of least action. The directional properties of the vectors <strong>in</strong> the Newtonian approach assist<br />

<strong>in</strong> our <strong>in</strong>tuition when sett<strong>in</strong>g up a problem, but the Lagrangian method is simpler mathematically when the<br />

mechanical system is more complex.<br />

The major advantage of the variational approaches to mechanics is that solution of the dynamical equations<br />

of motion can be simplified by express<strong>in</strong>g the motion <strong>in</strong> terms of <strong>in</strong>dependent generalized coord<strong>in</strong>ates.<br />

For Lagrangian mechanics these generalized coord<strong>in</strong>ates can be any set of <strong>in</strong>dependent variables,<br />

, where 1 ≤ ≤ , plus the correspond<strong>in</strong>g velocities ˙ . These <strong>in</strong>dependent generalized coord<strong>in</strong>ates<br />

completely specify the scalar potential and k<strong>in</strong>etic energies used <strong>in</strong> the Lagrangian or Hamiltonian. The<br />

variational approach allows for a much larger arsenal of possible generalized coord<strong>in</strong>ates than the typical<br />

vector coord<strong>in</strong>ates used <strong>in</strong> Newtonian mechanics. For example, the generalized coord<strong>in</strong>ates can be dimensionless<br />

amplitudes for the normal modes of coupled oscillator systems, or action-angle variables. Moreover,<br />

very different generalized coord<strong>in</strong>ates can be used for each of the variables. The tremendous freedom<br />

plus flexibility of the choice of generalized coord<strong>in</strong>ates is important when constra<strong>in</strong>t forces are act<strong>in</strong>g on the<br />

system. Generalized coord<strong>in</strong>ates allow the constra<strong>in</strong>t forces to be ignored by <strong>in</strong>clud<strong>in</strong>g auxiliary conditions<br />

to account for the k<strong>in</strong>ematic constra<strong>in</strong>ts that lead to correlated motion. The Lagrange method provides<br />

an <strong>in</strong>credibly consistent and mechanistic problem-solv<strong>in</strong>g strategy for many-body systems subject to constra<strong>in</strong>ts.<br />

Expressed <strong>in</strong> terms of generalized coord<strong>in</strong>ates, the Lagrange’s equations can be applied to a wide<br />

variety of physical problems <strong>in</strong>clud<strong>in</strong>g those <strong>in</strong>volv<strong>in</strong>g fields. The manipulation of scalar quantities <strong>in</strong> a<br />

configuration space of generalized coord<strong>in</strong>ates can greatly simplify problems compared with be<strong>in</strong>g conf<strong>in</strong>ed<br />

to a rigid orthogonal coord<strong>in</strong>ate system characterized by the Newtonian vector approach.<br />

The use of generalized coord<strong>in</strong>ates <strong>in</strong> Lagrange’s equations of motion can be applied to a wide range<br />

of physical phenomena <strong>in</strong>clud<strong>in</strong>g field theory, such as for electromagnetic fields, which are beyond the applicability<br />

of Newton’s equations of motion. The superiority of the Lagrangian approach compared to the<br />

Newtonian approach for solv<strong>in</strong>g problems <strong>in</strong> mechanics is apparent when deal<strong>in</strong>g with holonomic constra<strong>in</strong>t<br />

forces. Constra<strong>in</strong>t forces must be known and <strong>in</strong>cluded explicitly <strong>in</strong> the Newtonian equations of motion. Unfortunately,<br />

knowledge of the equations of motion is required to derive these constra<strong>in</strong>t forces. For holonomic<br />

constra<strong>in</strong>ed systems, the equations of motion can be solved directly without calculat<strong>in</strong>g the constra<strong>in</strong>t forces<br />

us<strong>in</strong>g the m<strong>in</strong>imal set of generalized coord<strong>in</strong>ate approach to Lagrangian mechanics. Moreover, the Lagrange<br />

approach has significant philosophical advantages compared to the Newtonian approach.


6.14. SUMMARY 169<br />

6.14 Summary<br />

Newtonian plausibility argument for Lagrangian mechanics:<br />

Ajustification for <strong>in</strong>troduc<strong>in</strong>g the calculus of variations to classical mechanics becomes apparent when<br />

the concept of the Lagrangian ≡ − is used <strong>in</strong> the functional and time is the <strong>in</strong>dependent variable.<br />

It was shown that Newton’s equation of motion can be rewritten as<br />

<br />

− = <br />

˙ <br />

(612)<br />

where <br />

<br />

are the excluded forces of constra<strong>in</strong>t plus any other conservative or non-conservative forces not<br />

<strong>in</strong>cluded <strong>in</strong> the potential This corresponds to the Euler-Lagrange equation for determ<strong>in</strong><strong>in</strong>g the m<strong>in</strong>imum<br />

of the time <strong>in</strong>tegral of the Lagrangian. Equation 612 can be written as<br />

<br />

− =<br />

˙ <br />

X<br />

<br />

() <br />

+ <br />

<br />

(615)<br />

where the Lagrange multiplier term accounts for holonomic constra<strong>in</strong>t forces, and <br />

<br />

<strong>in</strong>cludes all ad-<br />

. The<br />

of equation<br />

ditional forces not accounted for by the scalar potential , or the Lagrange multiplier terms <br />

<br />

constra<strong>in</strong>t forces can be <strong>in</strong>cluded explicitly as generalized forces <strong>in</strong> the excluded term <br />

<br />

615.<br />

d’Alembert’s Pr<strong>in</strong>ciple<br />

It was shown that d’Alembert’s Pr<strong>in</strong>ciple<br />

X<br />

(F − ṗ ) · r =0 (625)<br />

<br />

cleverly transforms the pr<strong>in</strong>ciple of virtual work from the realm of statics to dynamics. Application of virtual<br />

work to statics primarily leads to algebraic equations between the forces, whereas d’Alembert’s pr<strong>in</strong>ciple<br />

applied to dynamics leads to differential equations of motion.<br />

Lagrange equations from d’Alembert’s Pr<strong>in</strong>ciple<br />

After transform<strong>in</strong>g to generalized coord<strong>in</strong>ates, d’Alembert’s Pr<strong>in</strong>ciple leads to<br />

X<br />

∙½ <br />

<br />

<br />

µ <br />

˙ <br />

<br />

− <br />

<br />

¾<br />

− ¸<br />

=0 (638)<br />

If all the generalized coord<strong>in</strong>ates are <strong>in</strong>dependent, then equation 638 implies that the term <strong>in</strong> the square<br />

brackets is zero for each <strong>in</strong>dividual value of . That is, this implies the basic Euler-Lagrange equations of<br />

motion.<br />

The handl<strong>in</strong>g of both conservative and non-conservative generalized forces is best achieved by assum<strong>in</strong>g<br />

that the generalized force = P <br />

F · ¯r<br />

<br />

can be partitioned <strong>in</strong>to a conservative velocity-<strong>in</strong>dependent term,<br />

that can be expressed <strong>in</strong> terms of the gradient of a scalar potential, −∇ plus an excluded generalized force<br />

<br />

which conta<strong>in</strong>s the non-conservative, velocity-dependent, and all the constra<strong>in</strong>t forces not explicitly<br />

<strong>in</strong>cluded <strong>in</strong> the potential .Thatis,<br />

= −∇ + <br />

<br />

(641)<br />

Insert<strong>in</strong>g (641) <strong>in</strong>to (638) and assum<strong>in</strong>g that the potential is velocity <strong>in</strong>dependent, allows(638) to be<br />

rewritten as<br />

X<br />

∙½ µ <br />

¾ ¸<br />

( − ) ( − )<br />

− − <br />

=0<br />

(642)<br />

˙ <br />

<br />

Expressed <strong>in</strong> terms of the standard Lagrangian = − this gives<br />

X<br />

∙½ <br />

<br />

<br />

µ <br />

− ¾ ¸<br />

− <br />

=0<br />

˙ <br />

(644)


170 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

Note that equation (644) conta<strong>in</strong>s the basic Euler-Lagrange equation (638) for the special case when<br />

=0. In addition, note that if all the generalized coord<strong>in</strong>ates are <strong>in</strong>dependent, then the square bracket<br />

terms are zero for each value of which leads to the general Euler-Lagrange equations of motion<br />

½ µ <br />

− ¾<br />

= <br />

<br />

(645)<br />

˙ <br />

where ≥ ≥ 1.<br />

Newtonian mechanics has trouble handl<strong>in</strong>g constra<strong>in</strong>t forces because they lead to coupl<strong>in</strong>g of the degrees<br />

of freedom. Lagrangian mechanics is more powerful s<strong>in</strong>ce it provides the follow<strong>in</strong>g three ways to handle such<br />

correlated motion.<br />

1) M<strong>in</strong>imal set of generalized coord<strong>in</strong>ates<br />

If the coord<strong>in</strong>ates are <strong>in</strong>dependent, then the square bracket equals zero for each value of <strong>in</strong> equation<br />

644, which corresponds to Euler’s equation for each of the <strong>in</strong>dependent coord<strong>in</strong>ates. If the generalized<br />

coord<strong>in</strong>ates are coupled by constra<strong>in</strong>ts, then the coord<strong>in</strong>ates can be transformed to a m<strong>in</strong>imal set of<br />

= − <strong>in</strong>dependent coord<strong>in</strong>ates which then can be solved by apply<strong>in</strong>g equation 645 to the m<strong>in</strong>imal set<br />

of <strong>in</strong>dependent coord<strong>in</strong>ates.<br />

2) Lagrange multipliers approach<br />

The Lagrangian method concentrates solely on active forces, completely ignor<strong>in</strong>g all other <strong>in</strong>ternal forces.<br />

In Lagrangian mechanics the generalized forces, correspond<strong>in</strong>g to each generalized coord<strong>in</strong>ate, can be partitioned<br />

three ways<br />

= −∇ +<br />

X<br />

=1<br />

<br />

(q)+ <br />

<br />

<br />

where the velocity-<strong>in</strong>dependent conservative forces can be absorbed <strong>in</strong>to a scalar potential , the holonomic<br />

constra<strong>in</strong>t forces can be handled us<strong>in</strong>g the Lagrange multiplier term P <br />

=1 <br />

<br />

(q), and the rema<strong>in</strong><strong>in</strong>g<br />

part of the active forces can be absorbed <strong>in</strong>to the generalized force <br />

. The scalar potential energy is<br />

handled by absorb<strong>in</strong>g it <strong>in</strong>to the standard Lagrangian = −. If the constra<strong>in</strong>t forces are holonomic then<br />

these forces are easily and elegantly handled by use of Lagrange multipliers. All rema<strong>in</strong><strong>in</strong>g forces, <strong>in</strong>clud<strong>in</strong>g<br />

dissipative forces, can be handled by <strong>in</strong>clud<strong>in</strong>g them explicitly <strong>in</strong> the the generalized force <br />

.<br />

Comb<strong>in</strong><strong>in</strong>g the above two equations gives<br />

"<br />

X ½ µ <br />

− ¾<br />

#<br />

X<br />

− <br />

− (q) =0<br />

(656)<br />

˙ <br />

<br />

Use of the Lagrange multipliers to handle the constra<strong>in</strong>t forces ensures that all <strong>in</strong>f<strong>in</strong>itessimals are<br />

<strong>in</strong>dependent imply<strong>in</strong>g that the expression <strong>in</strong> the square bracket must be zero for each of the values of .<br />

This leads to Lagrange equations plus constra<strong>in</strong>t relations<br />

½ <br />

<br />

=1<br />

µ <br />

− ¾<br />

= <br />

+<br />

˙ <br />

X<br />

=1<br />

<br />

<br />

<br />

(q)<br />

(660)<br />

where =1 2 3 <br />

3) Generalized forces approach<br />

The two right-hand terms <strong>in</strong> (660) can be understood to be those forces act<strong>in</strong>g on the system that are<br />

not<br />

P<br />

absorbed <strong>in</strong>to the scalar potential component of the Lagrangian . The Lagrange multiplier terms<br />

<br />

=1 <br />

<br />

(q) account for the holonomic forces of constra<strong>in</strong>t that are not <strong>in</strong>cluded <strong>in</strong> the conservative<br />

potential or <strong>in</strong> the generalized forces <br />

. The generalized force<br />

<br />

=<br />

X<br />

<br />

F <br />

· r <br />

<br />

(617)<br />

is the sum of the components <strong>in</strong> the direction for all external forces that have not been taken <strong>in</strong>to account<br />

by the scalar potential or the Lagrange multipliers. Thus the non-conservative generalized force <br />

<br />

conta<strong>in</strong>s non-holonomic constra<strong>in</strong>t forces, <strong>in</strong>clud<strong>in</strong>g dissipative forces such as drag or friction, that are not<br />

<strong>in</strong>cluded <strong>in</strong> or used <strong>in</strong> the Lagrange multiplier terms to account for the holonomic constra<strong>in</strong>t forces.


6.14. SUMMARY 171<br />

Apply<strong>in</strong>g the Euler-Lagrange equations <strong>in</strong> mechanics:<br />

The optimal way to exploit Lagrangian mechanics is as follows:<br />

1. Select a set of <strong>in</strong>dependent generalized coord<strong>in</strong>ates.<br />

2. Partition the active forces <strong>in</strong>to three groups:<br />

(a) Conservative one-body forces<br />

(b) Holonomic constra<strong>in</strong>t forces<br />

(c) Generalized forces<br />

3. M<strong>in</strong>imize the number of generalized coord<strong>in</strong>ates.<br />

4. Derive the Lagrangian<br />

5. Derive the equations of motion<br />

Velocity-dependent Lorentz force:<br />

Usually velocity-dependent forces are non-holonomic. However, electromagnetism is a special case where<br />

the velocity-dependent Lorentz force F = (E + v × B) can be obta<strong>in</strong>ed from a velocity-dependent potential<br />

function ( ). Itwasshownthatthevelocity-dependentpotential<br />

= Φ − v · A<br />

(674)<br />

leads to the Lorentz force where Φ is the scalar electric potential and A the vector potential.<br />

Time-dependent forces:<br />

It was shown that time-dependent forces can lead to complicated motion hav<strong>in</strong>g both stable regions and<br />

unstable regions of motion that can exhibit chaos.<br />

Impulsive forces:<br />

A generalized impulse ˜ can be derived for an <strong>in</strong>stantaneous impulsive force from the time <strong>in</strong>tegral of<br />

the impulsive forces P given by equation 2135 us<strong>in</strong>g the time <strong>in</strong>tegral of equation 678, thatis<br />

Z +<br />

∆ = ˜ = lim<br />

→0<br />

<br />

Z +<br />

<br />

≡ lim<br />

→0<br />

<br />

X<br />

<br />

F · r <br />

<br />

= X <br />

˜P · r <br />

<br />

(679)<br />

Note that the generalized impulse ˜ can be a translational impulse ˜P with correspond<strong>in</strong>g translational<br />

variable or an angular impulsive torque ˜T with correspond<strong>in</strong>g angular variable .<br />

Comparison of Newtonian and Lagrangian mechanics:<br />

In contrast to Newtonian mechanics, which is based on know<strong>in</strong>g all the vector forces act<strong>in</strong>g on a system,<br />

Lagrangian mechanics can derive the equations of motion us<strong>in</strong>g generalized coord<strong>in</strong>ates without requir<strong>in</strong>g<br />

knowledge of the constra<strong>in</strong>t forces act<strong>in</strong>g on the system. Lagrangian mechanics provides a remarkably<br />

powerful, and <strong>in</strong>credibly consistent approach to solv<strong>in</strong>g for the equations of motion <strong>in</strong> classical mechanics,<br />

and is especially powerful for handl<strong>in</strong>g systems that are subject to holonomic constra<strong>in</strong>ts.


172 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

Problems<br />

1. Adiskofmass and radius rolls without slipp<strong>in</strong>g down a plane <strong>in</strong>cl<strong>in</strong>ed from the horizontal by an angle<br />

. The disk has a short weightless axle of negligible radius. From this axis is suspended a simple pendulum of<br />

length and whose bob has a mass . Assume that the motion of the pendulum takes place <strong>in</strong> the plane<br />

of the disk.<br />

(a) What generalized coord<strong>in</strong>ates would be appropriate for this situation?<br />

(b) Are there any equations of constra<strong>in</strong>t? If so, what are they?<br />

(c) F<strong>in</strong>d Lagrange’s equations for this system.<br />

2. A Lagrangian for a particular system can be written as<br />

= 2 ( ˙2 +2 ˙ ˙ + ˙ 2 ) − 2 (2 +2 + 2 )<br />

where and are arbitrary constants, but subject to the condition that 2 − 4 6= 0.<br />

(a) What are the equations of motion?<br />

(b) Exam<strong>in</strong>e the case =0=. What physical system does this represent?<br />

(c) Exam<strong>in</strong>e the case =0and = −. What physical system does this represent?<br />

(d) Based on your answers to (b) and (c), determ<strong>in</strong>e the physical system represented by the Lagrangian given<br />

above.<br />

3. Consider a particle of mass mov<strong>in</strong>g <strong>in</strong> a plane and subject to an <strong>in</strong>verse square attractive force.<br />

(a) Obta<strong>in</strong> the equations of motion.<br />

(b) Is the angular momentum about the orig<strong>in</strong> conserved?<br />

(c) Obta<strong>in</strong> expressions for the generalized forces. Recall that the generalized forces are def<strong>in</strong>ed by<br />

= X <br />

<br />

<br />

<br />

<br />

4. Consider a Lagrangian function of the form ( ˙<br />

¨ ). Here the Lagrangian conta<strong>in</strong>s a time derivative<br />

of the generalized coord<strong>in</strong>ates that is higher than the first. When work<strong>in</strong>g with such Lagrangians, the term<br />

“generalized mechanics” is used.<br />

(a) Consider a system with one degree of freedom. By apply<strong>in</strong>g the methods of the calculus of variations,<br />

and assum<strong>in</strong>g that Hamilton’s pr<strong>in</strong>ciple holds with respect to variations which keep both and ˙ fixed at<br />

the end po<strong>in</strong>ts, show that the correspond<strong>in</strong>g Lagrange equation is<br />

2 µ <br />

2 ¨<br />

<br />

− <br />

µ <br />

˙<br />

<br />

+ <br />

=0<br />

Such equations of motion have <strong>in</strong>terest<strong>in</strong>g applications <strong>in</strong> chaos theory.<br />

(b) Apply this result to the Lagrangian<br />

Do you recognize the equations of motion?<br />

= − 2 ¨ − 2 2 <br />

5. A bead of mass slides under gravity along a smooth wire bent <strong>in</strong> the shape of a parabola 2 = <strong>in</strong> the<br />

vertical ( ) plane.<br />

(a) What k<strong>in</strong>d (holonomic, nonholonomic, scleronomic, rheonomic) of constra<strong>in</strong>t acts on ?<br />

(b) Set up Lagrange’s equation of motion for with the constra<strong>in</strong>t embedded.


6.14. SUMMARY 173<br />

(c) Set up Lagrange’s equations of motion for both and with the constra<strong>in</strong>t adjo<strong>in</strong>ed and a Lagrangian<br />

multiplier <strong>in</strong>troduced.<br />

(d) Show that the same equation of motion for results from either of the methods used <strong>in</strong> part (b) or part<br />

(c).<br />

(e) Express <strong>in</strong> terms of and ˙.<br />

(f) What are the and components of the force of constra<strong>in</strong>t <strong>in</strong> terms of and ˙?<br />

6. Consider the two Lagrangians<br />

( ˙; ) and 0 ( ˙; ) =( ˙; )+<br />

( )<br />

<br />

where ( ) is an arbitrary function of the generalized coord<strong>in</strong>ates (). Show that these two Lagrangians<br />

yield the same Euler-Lagrange equations. As a consequence two Lagrangians that differ only by an exact time<br />

derivative are said to be equivalent.<br />

7. Consider the double pendulum compris<strong>in</strong>g masses 1 and 2 connected by <strong>in</strong>extensible str<strong>in</strong>gs as shown <strong>in</strong><br />

the figure. Assume that the motion of the pendulum takes place <strong>in</strong> a vertical plane.<br />

(a) Are there any equations of constra<strong>in</strong>t? If so, what are they?<br />

(b) F<strong>in</strong>d Lagrange’s equations for this system.<br />

O<br />

1<br />

L 1<br />

L 2<br />

m 1<br />

m 1 g<br />

2<br />

m 2<br />

m 2 g<br />

8 Consider the system shown <strong>in</strong> the figure which consists of a mass suspended via a constra<strong>in</strong>ed massless l<strong>in</strong>k<br />

of length where the po<strong>in</strong>t is acted upon by a spr<strong>in</strong>g of spr<strong>in</strong>g constant . The spr<strong>in</strong>g is unstretched when<br />

the massless l<strong>in</strong>k is horizontal. Assume that the holonomic constra<strong>in</strong>ts at and are frictionless.<br />

a Derive the equations of motion for the system us<strong>in</strong>g the method of Lagrange multipliers.<br />

kx<br />

x 0<br />

x<br />

L<br />

y<br />

mg<br />

9 Consider a pendulum, with mass , connected to a (horizontally) moveable support of mass .<br />

(a) Determ<strong>in</strong>e the Lagrangian of the system.<br />

(b) Determ<strong>in</strong>e the equations of motion for ¿ 1.<br />

(c) F<strong>in</strong>d an equation of motion <strong>in</strong> alone. What is the frequency of oscillation?<br />

(d) What is the frequency of oscillation for À ? Doesthismakesense?


174 CHAPTER 6. LAGRANGIAN DYNAMICS<br />

10 A sphere of radius is constra<strong>in</strong>ed to roll without slipp<strong>in</strong>g on the lower half of the <strong>in</strong>ner surface of a hollow<br />

cyl<strong>in</strong>der of radius Determ<strong>in</strong>e the Lagrangian function, the equation of constra<strong>in</strong>t, and the Lagrange equations<br />

of motion. F<strong>in</strong>d the frequency of small oscillations.<br />

11 A particle moves <strong>in</strong> a plane under the <strong>in</strong>fluence of a force = − −1 directed toward the orig<strong>in</strong>; and<br />

( 0) are constants. Choose generalized coord<strong>in</strong>ates with the potential energy zero at the orig<strong>in</strong>.<br />

a) F<strong>in</strong>d the Lagrangian equations of motion.<br />

b) Is the angular momentum about the orig<strong>in</strong> conserved?<br />

c) Is the total energy conserved?<br />

12 Two blocks, each of mass are connected by an extensionless, uniform str<strong>in</strong>g of length . One block is placed<br />

on a frictionless horizontal surface, and the other block hangs over the side, the str<strong>in</strong>g pass<strong>in</strong>g over a frictionless<br />

pulley. Describe the motion of the system:<br />

a) when the mass of the str<strong>in</strong>g is negligible<br />

b) when the str<strong>in</strong>g has mass .<br />

13 Two masses 1 and 2 ( 1 6= 2 ) are connected by a rigid rod of length and of negligible mass. An<br />

extensionless str<strong>in</strong>g of length 1 is attached to 1 and connected to a fixed po<strong>in</strong>t of the support . Similarly<br />

astr<strong>in</strong>goflength 2 ( 1 6= 2 ) connects 2 and . Obta<strong>in</strong> the equation of motion describ<strong>in</strong>g the motion <strong>in</strong><br />

the plane of 1 2 and ,andf<strong>in</strong>d the frequency of small oscillation around the equilibrium position.<br />

14 A th<strong>in</strong> uniform rigid rod of length 2 and mass is suspended by a massless str<strong>in</strong>g of length . Initiallythe<br />

system is hang<strong>in</strong>g vertically downwards <strong>in</strong> the gravitational field . Use as generalized coord<strong>in</strong>ates the angles<br />

given<strong>in</strong>thediagram.<br />

a) Derive the Lagrangian for the system.<br />

b) Use the Lagrangian to derive the equations of motion.<br />

c) A horizontal impulsive force <strong>in</strong> the direction strikes the bottom end of the rod for an <strong>in</strong>f<strong>in</strong>itessimal<br />

time . Derive the <strong>in</strong>itial conditions for the system immediately after the impulse has occurred.<br />

d) Draw a diagram show<strong>in</strong>g the geometry of the pendulum shortly after the impulse when the displacement<br />

angles are significant.<br />

O<br />

x<br />

1<br />

l<br />

y<br />

2<br />

2L<br />

Fx<br />

Mg


Chapter 7<br />

Symmetries, Invariance and the<br />

Hamiltonian<br />

7.1 Introduction<br />

The chapter 7 discussion of Lagrangian dynamics illustrates the power of Lagrangian mechanics for deriv<strong>in</strong>g<br />

the equations of motion. In contrast to Newtonian mechanics, which is expressed <strong>in</strong> terms of force vectors<br />

act<strong>in</strong>g on a system, the Lagrangian method, based on d’Alembert’s Pr<strong>in</strong>ciple or Hamilton’s Pr<strong>in</strong>ciple, is<br />

expressed <strong>in</strong> terms of the scalar k<strong>in</strong>etic and potential energies of the system. The Lagrangian approach is a<br />

sophisticated alternative to Newton’s laws of motion, that provides a simpler derivation of the equations of<br />

motion that allows constra<strong>in</strong>t forces to be ignored. In addition, the use of Lagrange multipliers or generalized<br />

forces allows the Lagrangian approach to determ<strong>in</strong>e the constra<strong>in</strong>t forces when these forces are of <strong>in</strong>terest.<br />

The equations of motion, derived either from Newton’s Laws or Lagrangian dynamics, can be non-trivial to<br />

solve mathematically. It is necessary to <strong>in</strong>tegrate second-order differential equations, which for degrees of<br />

freedom, imply 2 constants of <strong>in</strong>tegration.<br />

Chapter 7 will explore the remarkable connection between symmetry and <strong>in</strong>variance of a system under<br />

transformation, and the related conservation laws that imply the existence of constants of motion. Even<br />

when the equations of motion cannot be solved easily, it is possible to derive important physical pr<strong>in</strong>ciples<br />

regard<strong>in</strong>g the first-order <strong>in</strong>tegrals of motion of the system directly from the Lagrange equation, as well as for<br />

elucidat<strong>in</strong>g the underly<strong>in</strong>g symmetries plus <strong>in</strong>variance. This property is conta<strong>in</strong>ed <strong>in</strong> Noether’s theorem<br />

which states that conservation laws are associated with differentiable symmetries of a physical system.<br />

7.2 Generalized momentum<br />

Consider a holonomic system of masses under the <strong>in</strong>fluence of conservative forces that depend on position<br />

but not velocity ˙ , that is, the potential is velocity <strong>in</strong>dependent. Then for the coord<strong>in</strong>ate of particle <br />

for particles<br />

<br />

= − = <br />

(7.1)<br />

˙ ˙ ˙ ˙ <br />

=<br />

<br />

˙ <br />

<br />

X<br />

=1<br />

= ˙ = <br />

1<br />

2 ¡<br />

˙<br />

2<br />

+ ˙ 2 + ˙ 2 ¢<br />

Thus for a holonomic, conservative, velocity-<strong>in</strong>dependent potential we have<br />

which is the component of the l<strong>in</strong>ear momentum for the particle.<br />

<br />

˙ <br />

= (7.2)<br />

175


176 CHAPTER 7. SYMMETRIES, INVARIANCE AND THE HAMILTONIAN<br />

This result suggests an obvious extension of the concept of momentum to generalized coord<strong>in</strong>ates. The<br />

generalized momentum associated with the coord<strong>in</strong>ate is def<strong>in</strong>ed to be<br />

<br />

≡ (7.3)<br />

˙ <br />

Note that also is called the conjugate momentum or canonical momentum to where are<br />

conjugate, or canonical, variables. Remember that the l<strong>in</strong>ear momentum is the first-order time <strong>in</strong>tegral<br />

given by equation 210. If is not a spatial coord<strong>in</strong>ate, then is the generalized momentum, not the<br />

k<strong>in</strong>ematic l<strong>in</strong>ear momentum. For example, if is an angle, then will be angular momentum. That<br />

is, the generalized momentum may differ from the usual l<strong>in</strong>ear or angular momentum s<strong>in</strong>ce the def<strong>in</strong>ition<br />

(73) is more general than the usual = ˙ def<strong>in</strong>ition of l<strong>in</strong>ear momentum <strong>in</strong> classical mechanics. This is<br />

illustrated by the case of a mov<strong>in</strong>g charged particles <strong>in</strong> an electromagnetic field. Chapter 6 showed<br />

that electromagnetic forces on a charge can be described <strong>in</strong> terms of a scalar potential where<br />

= (Φ − A · v ) (7.4)<br />

Thus the Lagrangian for the electromagnetic force can be written as<br />

X<br />

∙ 1<br />

=<br />

2 v · v − (Φ − A · v )¸<br />

=1<br />

(7.5)<br />

The generalized momentum to the coord<strong>in</strong>ate for charge and mass is given by the above Lagrangian<br />

= <br />

˙ <br />

= ˙ + (7.6)<br />

Note that this <strong>in</strong>cludes both the mechanical l<strong>in</strong>ear momentum plus the correct electromagnetic momentum.<br />

The fact that the electromagnetic field carries momentum should not be a surprise s<strong>in</strong>ce electromagnetic<br />

waves also carry energy as is illustrated by the transmission of radiant energy from the sun.<br />

7.1 Example: Feynman’s angular-momentum paradox<br />

Feynman[Fey84] posed the follow<strong>in</strong>g paradox. A circular <strong>in</strong>sulat<strong>in</strong>g disk mounted on frictionless bear<strong>in</strong>gs,<br />

has a circular r<strong>in</strong>g of total charge uniformly distributed around the perimeter of the circular disk at the<br />

radius . A superconduct<strong>in</strong>g long solenoid of radius where , is fixedtothediskandismounted<br />

coaxial with the bear<strong>in</strong>gs. The moment of <strong>in</strong>ertia of the system about the rotation axis is . Initially the disk<br />

plus superconduct<strong>in</strong>g solenoid are stationary with a steady current produc<strong>in</strong>g a uniform magnetic field 0<br />

<strong>in</strong>side the solenoid. Assume that a rise <strong>in</strong> temperature of the solenoid destroys the superconductivity lead<strong>in</strong>g<br />

to a rapid dissipation of the electric current and resultant magnetic field. Assume that the system is free to<br />

rotate, no other forces or torques are act<strong>in</strong>g on the system, and that the charge carriers <strong>in</strong> the solenoid have<br />

zero mass and thus do not contribute to the angular momentum. Does the system rotate when the current <strong>in</strong><br />

the solenoid stops?<br />

Initially the system is stationary with zero mechanical angular<br />

momentum. Faraday’s Law states that, when the magnetic<br />

field dissipates from 0 to zero, there will be a torque N act<strong>in</strong>g<br />

on the circumferential charge at radius due to the change<br />

<strong>in</strong> magnetic flux Φ.<br />

N() =− Φ<br />

<br />

S<strong>in</strong>ce Φ<br />

<br />

0, this torque leads to an angular impulse which<br />

will equal the f<strong>in</strong>al mechanical angular momentum.<br />

Z<br />

L <br />

= T = N() = Φ


7.3. INVARIANT TRANSFORMATIONS AND NOETHER’S THEOREM 177<br />

The <strong>in</strong>itial angular momentum <strong>in</strong> the electromagnetic field can be derived us<strong>in</strong>g equation 76 plus Stoke’s<br />

theorem, Equation 2142 gives that the f<strong>in</strong>al angular momentum equals the angular impulse<br />

Z I<br />

I<br />

I<br />

Z<br />

L <br />

= ˙ = = = B · dS =Φ<br />

<br />

I Z<br />

where Φ = = B · dS is the <strong>in</strong>itial total magnetic flux through the solenoid. Thus the total <strong>in</strong>itial<br />

angular momentum is given by<br />

L <br />

=0+L <br />

= Φ<br />

S<strong>in</strong>ce the f<strong>in</strong>al electromagnetic field is zero the f<strong>in</strong>al total angular momentum is given by<br />

L <br />

<br />

= L <br />

<br />

+0=Φ<br />

Note that the total angular momentum is conserved. That is, <strong>in</strong>itially all the angular momentum is stored <strong>in</strong><br />

the electromagnetic field, whereas the f<strong>in</strong>al angular momentum is all mechanical. This expla<strong>in</strong>s the paradox<br />

that the mechanical angular momentum is not conserved, only the total angular momentum of the system is<br />

conserved, that is, the sum of the mechanical and electromagnetic angular momenta.<br />

7.3 Invariant transformations and Noether’s Theorem<br />

One of the great advantages of Lagrangian mechanics is the freedom it allows <strong>in</strong> choice of generalized<br />

coord<strong>in</strong>ates which can simplify derivation of the equations of motion. For example, for any set of coord<strong>in</strong>ates,<br />

a reversible po<strong>in</strong>t transformation can def<strong>in</strong>e another set of coord<strong>in</strong>ates 0 such that<br />

0 = 0 ( 1 2 ; ) (7.7)<br />

The new set of generalized coord<strong>in</strong>ates satisfies Lagrange’s equations of motion with the new Lagrangian<br />

( 0 ˙ 0 )=( ˙ ) (7.8)<br />

The Lagrangian is a scalar, with units of energy, which does not change if the coord<strong>in</strong>ate representation<br />

is changed. Thus ( 0 ˙ 0 ) can be derived from ( ˙ ) by substitut<strong>in</strong>g the <strong>in</strong>verse relation =<br />

(1 0 2 0 ; 0 ) <strong>in</strong>to ( ˙ ) That is, the value of the Lagrangian is <strong>in</strong>dependent of which coord<strong>in</strong>ate<br />

representation is used. Although the general form of Lagrange’s equations of motion is preserved <strong>in</strong> any<br />

po<strong>in</strong>t transformation, the explicit equations of motion for the new variables usually look different from those<br />

with the old variables. A typical example is the transformation from cartesian to spherical coord<strong>in</strong>ates. For<br />

a given system, there can be particular transformations for which the explicit equations of motion are the<br />

same for both the old and new variables. Transformations for which the equations of motion are <strong>in</strong>variant,<br />

are called <strong>in</strong>variant transformations. It will be shown that if the Lagrangian does not explicitly conta<strong>in</strong><br />

a particular coord<strong>in</strong>ate of displacement then the correspond<strong>in</strong>g conjugate momentum, is conserved.<br />

This relation is called Noether’s theorem which states “For each symmetry of the Lagrangian, there is a<br />

conserved quantity”.<br />

Noether’s Theorem will be used to consider <strong>in</strong>variant transformations for two dependent variables, ()<br />

and () plus their conjugate momenta and . For a closed system, these provide up to six possible<br />

conservation laws for the three axes. Then we will discuss the <strong>in</strong>dependent variable and its relation to<br />

the Generalized Energy Theorem, which provides another possible conservation law. For simplicity, these<br />

discussions will assume that the systems are holonomic and conservative.<br />

The Lagrange equations us<strong>in</strong>g generalized coord<strong>in</strong>ates for holonomic systems, was given by equation 660<br />

to be<br />

½ µ <br />

− ¾ X<br />

=<br />

˙ <br />

=1<br />

This can be written <strong>in</strong> terms of the generalized momentum as<br />

½ <br />

− ¾ X<br />

=<br />

<br />

=1<br />

<br />

<br />

<br />

(q)+ <br />

(7.9)<br />

<br />

<br />

<br />

(q)+ <br />

(7.10)


178 CHAPTER 7. SYMMETRIES, INVARIANCE AND THE HAMILTONIAN<br />

or equivalently as<br />

"<br />

˙ = <br />

<br />

#<br />

X <br />

+ (q)+ <br />

<br />

(7.11)<br />

<br />

=1<br />

Note that if the Lagrangian does not conta<strong>in</strong> explicitly, that is, the Lagrangian is <strong>in</strong>variant to a l<strong>in</strong>ear<br />

translation, or equivalently, is spatially homogeneous, and if the Lagrange multiplier constra<strong>in</strong>t force and<br />

generalized force terms are zero, then<br />

<br />

<br />

+<br />

" <br />

X<br />

=1<br />

In this case the Lagrange equation reduces to<br />

<br />

(q)+ <br />

<br />

<br />

#<br />

=0 (7.12)<br />

˙ = <br />

<br />

=0 (7.13)<br />

Equation 713 corresponds to be<strong>in</strong>g a constant of motion. Stated <strong>in</strong> words, the generalized momentum <br />

is a constant of motion if the Lagrangian is <strong>in</strong>variant to a spatial translation of , and the constra<strong>in</strong>t plus<br />

generalized force terms are zero. Expressed another way, if the Lagrangian does not conta<strong>in</strong> a given coord<strong>in</strong>ate<br />

and the correspond<strong>in</strong>g constra<strong>in</strong>t plus generalized forces are zero, then the generalized momentum<br />

associated with this coord<strong>in</strong>ate is conserved. Note that this example of Noether’s theorem applies to any<br />

component of q. For example, <strong>in</strong> the uniform gravitational field at the surface of the earth, the Lagrangian<br />

does not depend on the and coord<strong>in</strong>ates <strong>in</strong> the horizontal plane, thus and are conserved, whereas,<br />

due to the gravitational force, the Lagrangian does depend on the vertical axis and thus is not conserved.<br />

7.2 Example: Atwoods mach<strong>in</strong>e<br />

Assume that the l<strong>in</strong>ear momentum is conserved for the Atwood’s mach<strong>in</strong>e shown <strong>in</strong> the figure below. Let<br />

theleftmassriseadistance and the right mass rise a distance . Then the middle mass must drop by<br />

+ to conserve the length of the str<strong>in</strong>g. The Lagrangian of the system is<br />

= 1 2 (4) ˙2 + 1 2 (3)(− ˙− ˙)2 + 1 2 ˙2 −(4 +3(− − )+) = 7 2 ˙2 +3 ˙ ˙+2 ˙ 2 −(−2)<br />

Note that the transformation<br />

= 0 +2<br />

= 0 + <br />

results <strong>in</strong> the potential energy term (−2) =( 0 −2 0 )<br />

which is a constant of motion. As a result the Lagrangian<br />

is <strong>in</strong>dependent of which means that it is <strong>in</strong>variant to the<br />

small perturbation and thus <br />

<br />

=0 Therefore, accord<strong>in</strong>g<br />

to Noether’s theorem, the correspond<strong>in</strong>g l<strong>in</strong>ear momentum<br />

= <br />

˙<br />

is conserved. This conserved l<strong>in</strong>ear momentum<br />

then is given by<br />

x<br />

4m<br />

3m<br />

Example of an Atwood’s mach<strong>in</strong>e<br />

= <br />

˙ = ˙<br />

˙ ˙ + ˙<br />

= (7 ˙ +3˙)(2) + (3 ˙ +4˙) =(17 ˙ +10˙)<br />

˙ ˙<br />

Thus, if the system starts at rest with =0,then ˙ always equals − 10<br />

17 ˙ s<strong>in</strong>ce is constant.<br />

Note that this also can be shown us<strong>in</strong>g the Euler-Lagrange equations <strong>in</strong> that Λ =0and Λ =0give<br />

Add<strong>in</strong>g the second equation to twice the first gives<br />

7¨ +3¨ = −<br />

3¨ +4¨ = 2<br />

17¨ +10¨ = (17 ˙ +10 ˙) =0<br />

<br />

This is the result obta<strong>in</strong>ed directly us<strong>in</strong>g Noether’s theorem.<br />

m<br />

y


7.4. ROTATIONAL INVARIANCE AND CONSERVATION OF ANGULAR MOMENTUM 179<br />

7.4 Rotational <strong>in</strong>variance and conservation of angular momentum<br />

The arguments, used above, apply equally well to conjugate momenta and for rotation about any axis.<br />

The Lagrange equation is<br />

½ <br />

− ¾ X <br />

= <br />

<br />

(q)+ (7.14)<br />

=1<br />

If no constra<strong>in</strong>t or generalized torques act on the system, then the right-hand side of equation 714 is zero.<br />

Moreover if the Lagrangian <strong>in</strong> not an explicit function of then <br />

<br />

=0 and assum<strong>in</strong>g that the constra<strong>in</strong>t<br />

plus generalized torques are zero, then is a constant of motion.<br />

Noether’s Theorem illustrates this general result which can be stated as, if the Lagrangian is rotationally<br />

<strong>in</strong>variant about some axis, then the component of the angular momentum along that axis is conserved. Also<br />

this is true for the more general case where the Lagrangian is <strong>in</strong>variant to rotation about any axis, which<br />

leads to conservation of the total angular momentum.<br />

7.3 Example: Conservation of angular momentum for rotational <strong>in</strong>variance:<br />

The Noether theorem result for rotational-<strong>in</strong>variance about an<br />

axis also can be derived us<strong>in</strong>g cartesian coord<strong>in</strong>ates as shown below.<br />

As discussed <strong>in</strong> appendix , it is necessary to limit discussion of<br />

rotation to <strong>in</strong>f<strong>in</strong>itessimal rotation angles <strong>in</strong> order to represent the<br />

rotation by a vector. Consider an <strong>in</strong>f<strong>in</strong>itessimal rotation about<br />

some axis, which is a vector. As illustrated <strong>in</strong> the adjacent figure,<br />

this can be expressed as<br />

r = θ × r<br />

The velocity vectors also change on rotation of the system obey<strong>in</strong>g<br />

the transformation equation which is common to all vectors, that<br />

is,<br />

r<br />

ṙ = θ × ṙ<br />

If the Lagrangian is unaffected by the orientation of the system,<br />

that is, it is rotationally <strong>in</strong>variant, then it can be shown that the<br />

angular momentum is conserved. For example, consider that the<br />

Lagrangian is <strong>in</strong>variant to rotation about some axis . S<strong>in</strong>ce the<br />

Inf<strong>in</strong>itessimal rotation<br />

Lagrangian is a function<br />

= ( ˙ ; )<br />

then the expression that the Lagrangian does not change due to an <strong>in</strong>f<strong>in</strong>itesimal rotation about this axis<br />

can be expressed as<br />

= X <br />

where cartesian coord<strong>in</strong>ates have been used.<br />

Us<strong>in</strong>g the generalized momentum<br />

then, Lagrange’s equation gives<br />

that is<br />

Insert<strong>in</strong>g this <strong>in</strong>to equation gives<br />

<br />

<br />

+ X <br />

<br />

˙ <br />

= <br />

<br />

− <br />

<br />

=0<br />

˙ = <br />

<br />

<br />

˙ <br />

˙ =0<br />

()<br />

=<br />

3X<br />

˙ +<br />

<br />

3X<br />

˙ =0


180 CHAPTER 7. SYMMETRIES, INVARIANCE AND THE HAMILTONIAN<br />

This is equivalent to the scalar products<br />

ṗ · r + p · ṙ =0<br />

For an <strong>in</strong>f<strong>in</strong>itessimal rotation then = × and ˙ = × ˙ . Therefore<br />

Thecyclicordercanbepermutedgiv<strong>in</strong>g<br />

ṗ · (θ × r)+p · (θ × ṙ) =0<br />

θ · (r × ṗ)+θ · (ṙ × p) = 0<br />

θ · [(r × ṗ)+(ṙ × p)] = 0<br />

θ · (r × p)<br />

<br />

= 0<br />

Because the <strong>in</strong>f<strong>in</strong>itessimal angle is arbitrary, then the time derivative<br />

<br />

(r × p) =0<br />

<br />

about the axis of rotation But the bracket (r × p) equals the angular momentum. That is;<br />

Angular momentum = (r × p) =constant<br />

This proves the Noether’ theorem that the angular momentum about any axis is conserved if the Lagrangian<br />

is rotationally <strong>in</strong>variant about that axis.<br />

7.4 Example: Diatomic molecules and axially-symmetric nuclei<br />

An <strong>in</strong>terest<strong>in</strong>g example of Noether’s theorem applies to diatomic molecules such as 2 2 2 2 2<br />

and 2 . The electric field produced by the two charged nuclei of the diatomic molecule has cyl<strong>in</strong>drical<br />

symmetry about the axis through the two nuclei. Electrons are bound to this dumbbell arrangement of the two<br />

nuclear charges which may be rotat<strong>in</strong>g and vibrat<strong>in</strong>g <strong>in</strong> free space. Assum<strong>in</strong>g that there are no external torques<br />

act<strong>in</strong>g on the diatomic molecule <strong>in</strong> free space, then the angular momentum about any fixed axis <strong>in</strong> free space<br />

must be conserved accord<strong>in</strong>g to Noether’s theorem. If no external torques are applied, then the component of<br />

the angular momentum about any fixed axis is conserved, that is, the total angular momentum is conserved.<br />

What is especially <strong>in</strong>terest<strong>in</strong>g is that s<strong>in</strong>ce the electrostatic potential, and thus the Lagrangian, of the diatomic<br />

molecule has cyl<strong>in</strong>drical symmetry, that is <br />

<br />

=0, then the component of the angular momentum with respect<br />

to this symmetry axis also is conserved irrespective of how the diatomic molecule rotates or vibrates <strong>in</strong> free<br />

space. That is, an additional symmetry has been identified that leads to an additional conservation law that<br />

applies to the angular momentum.<br />

An example of Noether’s theorem is <strong>in</strong> nuclear physics where some nuclei have a spheroidal shape similar<br />

to an american football or a rugby ball. This spheroidal shape has an axis of symmetry along the long axis.<br />

The Lagrangian is rotationally <strong>in</strong>variant about the symmetry axis result<strong>in</strong>g <strong>in</strong> the angular momentum about<br />

the symmetry axis be<strong>in</strong>g conserved <strong>in</strong> addition to conservation of the total angular momentum.<br />

7.5 Cyclic coord<strong>in</strong>ates<br />

Translational and rotational <strong>in</strong>variance occurs when a system has a cyclic coord<strong>in</strong>ate Acycliccoord<strong>in</strong>ate<br />

is one that does not explicitly appear <strong>in</strong> the Lagrangian. The term cyclic is a natural name when one has<br />

cyl<strong>in</strong>drical or spherical symmetry. In Hamiltonian mechanics a cyclic coord<strong>in</strong>ate often is called an ignorable<br />

coord<strong>in</strong>ate. By virtue of Lagrange’s equations<br />

<br />

− =0 (7.15)<br />

˙ <br />

then a cyclic coord<strong>in</strong>ate is one for which <br />

<br />

=0.Thus<br />

<br />

= ˙ =0 (7.16)<br />

˙ <br />

that is, is a constant of motion if the conjugate coord<strong>in</strong>ate is cyclic. ThisisjustNoether’sTheorem.


7.6. KINETIC ENERGY IN GENERALIZED COORDINATES 181<br />

7.6 K<strong>in</strong>etic energy <strong>in</strong> generalized coord<strong>in</strong>ates<br />

Application of Noether’s theorem to the conservation of energy requires the k<strong>in</strong>etic energy to be expressed<br />

<strong>in</strong> generalized coord<strong>in</strong>ates. In terms of fixed rectangular coord<strong>in</strong>ates, the k<strong>in</strong>etic energy for bodies, each<br />

hav<strong>in</strong>g three degrees of freedom, is expressed as<br />

= 1 X 3X<br />

˙ 2 (7.17)<br />

2<br />

=1 =1<br />

These can be expressed <strong>in</strong> terms of generalized coord<strong>in</strong>ates as = ( ) and <strong>in</strong> terms of generalized<br />

velocities<br />

X <br />

˙ = ˙ + <br />

(7.18)<br />

<br />

=1 <br />

Tak<strong>in</strong>gthesquareof ˙ and <strong>in</strong>sert<strong>in</strong>g <strong>in</strong>to the k<strong>in</strong>etic energy relation gives<br />

(q ˙q)= X X 1<br />

2 <br />

<br />

˙ ˙ + X X <br />

˙ + X X<br />

µ 2<br />

1<br />

<br />

<br />

<br />

<br />

<br />

<br />

2 <br />

<br />

(7.19)<br />

<br />

<br />

This can be abbreviated as<br />

(q ˙q)= 2 (q ˙q)+ 1 (q ˙q)+ 0 (q) (7.20)<br />

where<br />

2 (q ˙q) = X X 1<br />

2 <br />

<br />

˙ ˙ = X ˙ ˙ (7.21)<br />

<br />

<br />

<br />

<br />

<br />

1 (q ˙q) = X X <br />

˙ = X ˙ (7.22)<br />

<br />

<br />

<br />

0 (q) = X X<br />

µ 2<br />

1<br />

2 <br />

<br />

(7.23)<br />

<br />

<br />

where<br />

X 3X 1<br />

≡<br />

2 <br />

<br />

(7.24)<br />

<br />

=1 =1<br />

<br />

When the transformed system is scleronomic, time does not appear explicitly <strong>in</strong> the transformation<br />

equations to generalized coord<strong>in</strong>ates s<strong>in</strong>ce <br />

<br />

=0. Then 1 = 0 =0, and the k<strong>in</strong>etic energy reduces to<br />

a homogeneous quadratic function of the generalized velocities<br />

(q ˙q)= 2 (q ˙q) (7.25)<br />

A useful relation can be derived by tak<strong>in</strong>g the differential of equation 721 with respect to ˙ .Thatis<br />

2 (q ˙q)<br />

= X ˙ + X ˙ (7.26)<br />

˙ <br />

<br />

<br />

Multiply this by ˙ and sum over gives<br />

X 2 (q ˙q)<br />

˙ = X ˙ ˙ + X ˙ ˙ =2 X ˙ ˙ =2 2<br />

˙ <br />

<br />

<br />

<br />

<br />

Similarly, the products of the generalized velocities ˙ with the correspond<strong>in</strong>g derivatives of 1 and 0 give<br />

X 2<br />

˙ = 2 2 (7.27)<br />

˙ <br />

<br />

X 1 (q ˙q)<br />

˙ = 1 (q ˙q) (7.28)<br />

˙ <br />

<br />

X 0 (q)<br />

˙ = 0 (7.29)<br />

˙


182 CHAPTER 7. SYMMETRIES, INVARIANCE AND THE HAMILTONIAN<br />

Equation 725 gives that = 2 when the transformed system is scleronomic, i.e. <br />

<br />

=0 andthenthe<br />

k<strong>in</strong>etic energy is a quadratic function of the generalized velocities ˙ .Us<strong>in</strong>gthedef<strong>in</strong>ition of the generalized<br />

momentum equation 73 assum<strong>in</strong>g = 2 , and that the potential is velocity <strong>in</strong>dependent, gives that<br />

≡ = − = 2<br />

(7.30)<br />

˙ ˙ ˙ ˙ <br />

Then equation 727 reduces to the useful relation that<br />

2 = 1 X<br />

˙ = 1 ˙q · p (7.31)<br />

2 2<br />

where, for compactness, the summation is abbreviated as a scalar product.<br />

7.7 Generalized energy and the Hamiltonian function<br />

<br />

Consider the time derivative of the Lagrangian, plus the fact that time is the <strong>in</strong>dependent variable <strong>in</strong> the<br />

Lagrangian. Then the total time derivative is<br />

<br />

= X <br />

<br />

<br />

˙ + X <br />

<br />

¨ + <br />

˙ <br />

(7.32)<br />

The Lagrange equations for a conservative force are given by equation 660 to be<br />

<br />

− <br />

X<br />

= <br />

+ (q) (7.33)<br />

˙ <br />

The holonomic constra<strong>in</strong>ts can be accounted for us<strong>in</strong>g the Lagrange multiplier terms while the generalized<br />

force <br />

<strong>in</strong>cludes non-holonomic forces or other forces not <strong>in</strong>cluded <strong>in</strong> the potential energy term of the<br />

Lagrangian, or holonomic forces not accounted for by the Lagrange multiplier terms.<br />

Substitut<strong>in</strong>g equation 733 <strong>in</strong>to equation 732 gives<br />

<br />

= X <br />

˙ − X "<br />

#<br />

X<br />

˙ <br />

+ (q) + X <br />

¨ + <br />

<br />

˙<br />

<br />

<br />

<br />

˙<br />

=1<br />

<br />

= X µ <br />

<br />

˙ − X "<br />

#<br />

X<br />

˙ <br />

+ (q) + <br />

(7.34)<br />

˙<br />

<br />

<br />

<br />

<br />

=1<br />

This can be written <strong>in</strong> the form<br />

⎡<br />

⎤<br />

<br />

⎣ X µ <br />

<br />

˙ − ⎦ = X "<br />

#<br />

X<br />

˙ <br />

+ (q) − <br />

(7.35)<br />

˙<br />

<br />

<br />

<br />

Def<strong>in</strong>e Jacobi’s Generalized Energy 1 (q ˙q) by<br />

(q ˙q) ≡ X µ <br />

<br />

˙ − (q ˙q) (7.36)<br />

˙<br />

<br />

Jacobi’s generalized momentum, equation 73 can be used to express the generalized energy ( ˙ ) <strong>in</strong><br />

terms of the canonical coord<strong>in</strong>ates ˙ and ,plustime. Def<strong>in</strong>e the Hamiltonian function to equal the<br />

generalized energy expressed <strong>in</strong> terms of the conjugate variables ( ),thatis,<br />

(q p) ≡ (q ˙q) ≡ X µ <br />

<br />

˙ − (q ˙q)= X ( ˙ ) − (q ˙q) (7.37)<br />

˙<br />

<br />

<br />

This Hamiltonian (q p) underlies Hamiltonian mechanics which plays a profoundly important role <strong>in</strong><br />

most branches of physics as illustrated <strong>in</strong> chapters 8 15 and 18.<br />

1 Most textbooks call the function (q ˙q) Jacobi’s energy <strong>in</strong>tegral. This book adopts the more descriptive name Generalized<br />

energy <strong>in</strong> analogy with use of generalized coord<strong>in</strong>ates q and generalized momentum p.<br />

=1<br />

=1


7.8. GENERALIZED ENERGY THEOREM 183<br />

7.8 Generalized energy theorem<br />

The Hamilton function, 737 plus equation 735 lead to the generalized energy theorem<br />

(q p) (q ˙q)<br />

= = X "<br />

#<br />

X<br />

˙ (q ˙q)<br />

+ (q) −<br />

<br />

<br />

<br />

<br />

<br />

=1<br />

h<br />

Note that for the special case where all the external forces <br />

+ P i<br />

<br />

=1 <br />

<br />

(q) =0,then<br />

(7.38)<br />

<br />

= −<br />

(7.39)<br />

<br />

h<br />

Thus the Hamiltonian is time <strong>in</strong>dependent if both <br />

+ P i<br />

<br />

=1 <br />

<br />

(q) =0and the Lagrangian are<br />

time-<strong>in</strong>dependent. For an isolated closed system hav<strong>in</strong>g no external forces act<strong>in</strong>g, then the Lagrangian is<br />

time <strong>in</strong>dependent because the velocities are constant, and there is no external potential energy. That is, the<br />

Lagrangian is time-<strong>in</strong>dependent, and<br />

<br />

<br />

⎡<br />

⎣ X <br />

µ <br />

<br />

˙ <br />

˙ <br />

⎤<br />

− ⎦ = = − =0 (7.40)<br />

<br />

As a consequence, the Hamiltonian (q p) and generalized energy (q ˙q), both are constants of motion<br />

if the Lagrangian is a constant of motion, and if the external non-potential forces are zero. This is an example<br />

of Noether’s theorem, where the symmetry of time <strong>in</strong>dependence leads to conservation of the conjugate<br />

variable, which is the Hamiltonian or Generalized energy.<br />

7.9 Generalized energy and total energy<br />

The generalized k<strong>in</strong>etic energy, equation 720, can be used to write the generalized Lagrangian as<br />

(q ˙q)= 2 (q ˙q)+ 1 (q ˙q)+ 0 (q) − (q) (7.41)<br />

If the potential energy does not depend explicitly on velocities ˙ or time, then<br />

= ( − )<br />

= = <br />

(7.42)<br />

˙ ˙ ˙ <br />

Equation 742 canbeusedtowritetheHamiltonian, equation737, as<br />

(q p) = X µ <br />

2<br />

˙ + X µ <br />

1<br />

˙ + X µ <br />

0<br />

˙ − (q ˙q) (7.43)<br />

˙<br />

<br />

˙<br />

<br />

˙<br />

<br />

<br />

Us<strong>in</strong>g equations 727 728 729 gives that the total generalized Hamiltonian (q p) equals<br />

(q p) =2 2 + 1 − ( 2 + 1 + 0 − ) = 2 − 0 + (7.44)<br />

But the sum of the k<strong>in</strong>etic and potential energies equals the total energy. Thus equation 744 can be rewritten<br />

<strong>in</strong> the form<br />

(q p) =( + ) − ( 1 +2 0 )= − ( 1 +2 0 ) (7.45)<br />

Note that Jacobi’s generalized energy and the Hamiltonian do not equal the total energy . However, <strong>in</strong><br />

the special case where the transformation is scleronomic, then 1 = 0 =0 and if the potential energy <br />

does not depend explicitly of ˙ , then the generalized energy (Hamiltonian) equals the total energy, that is,<br />

= Recognition of the relation between the Hamiltonian and the total energy facilitates determ<strong>in</strong><strong>in</strong>g<br />

the equations of motion.


184 CHAPTER 7. SYMMETRIES, INVARIANCE AND THE HAMILTONIAN<br />

7.10 Hamiltonian <strong>in</strong>variance<br />

Chapters 78 79 addressed two important and <strong>in</strong>dependent features of the Hamiltonian regard<strong>in</strong>g: ) when<br />

is conserved, and ) when equals the total mechanical energy. These important results are summarized<br />

below with a discussion of the assumptions made <strong>in</strong> deriv<strong>in</strong>g the Hamiltonian, as well as the implications.<br />

a) Conservation of generalized energy<br />

The generalized energy theorem (738) was given as<br />

(q p) (q ˙q)<br />

= = X "<br />

#<br />

X<br />

˙ (q ˙q)<br />

+ (q) −<br />

<br />

<br />

<br />

<br />

<br />

=1<br />

Note that when P h<br />

˙ <br />

+ P i<br />

<br />

=1 <br />

<br />

(q) =0,thenequation746 reduces to<br />

(7.46)<br />

<br />

= −<br />

(7.47)<br />

<br />

Also, when P h<br />

˙ <br />

+ P i<br />

<br />

=1 <br />

<br />

(q) =0 and if the Lagrangian is not an explicit function of time,<br />

then the Hamiltonian is a constant of motion. That is, is conserved if, and only if, the Lagrangian, and<br />

consequently the Hamiltonian, are not explicit functions of time, and if the external forces are zero.<br />

b) The generalized energy and total energy<br />

If the follow<strong>in</strong>g two requirements are satisfied<br />

1) The k<strong>in</strong>etic energy has a homogeneous quadratic dependence on the generalized velocities, that is, the<br />

transformation to generalized coord<strong>in</strong>ates is <strong>in</strong>dependent of time, <br />

<br />

=0<br />

2) The potential energy is not velocity dependent, thus the terms <br />

˙ <br />

=0<br />

Then equation 745 implies that the Hamiltonian equals the total mechanical energy, that is,<br />

= + = (7.48)<br />

Expressed <strong>in</strong> words, the generalized energy (Hamiltonian) equals the total energy if the constra<strong>in</strong>ts are<br />

time <strong>in</strong>dependent and the potential energy is velocity <strong>in</strong>dependent. This is equivalent to stat<strong>in</strong>g that, if the<br />

constra<strong>in</strong>ts, or generalized coord<strong>in</strong>ates, for the system are time <strong>in</strong>dependent, then = .<br />

The four comb<strong>in</strong>ations of the above two <strong>in</strong>dependent conditions, assum<strong>in</strong>g that the external forces term<br />

<strong>in</strong> equation 746 is zero, are summarized <strong>in</strong> table 71.<br />

Table 7.1: Hamiltonian and total energy<br />

Hamiltonian Constra<strong>in</strong>ts and coord<strong>in</strong>ate transformation<br />

Time behavior Time <strong>in</strong>dependent Time dependent<br />

<br />

<br />

= − <br />

<br />

=0 conserved, = conserved, 6= <br />

<br />

<br />

= − <br />

<br />

6=0 not conserved, = not conserved, 6= <br />

Note the follow<strong>in</strong>g general facts regard<strong>in</strong>g the Lagrangian and the Hamiltonian.<br />

(1) the Lagrangian is <strong>in</strong>def<strong>in</strong>ite with respect to addition of a constant to the scalar potential,<br />

(2) the Lagrangian is <strong>in</strong>def<strong>in</strong>ite with respect to addition of a constant velocity,<br />

(3) there is no unique choice of generalized coord<strong>in</strong>ates.<br />

(4) the Hamiltonian is a scalar function that is derived from the Lagrangian scalar function.<br />

(5) the generalized momentum is derived from the Lagrangian.<br />

These facts, plus the ability to recognize the conditions under which is conserved, and when = <br />

can greatly facilitate solv<strong>in</strong>g problems as shown by the follow<strong>in</strong>g two examples.


7.10. HAMILTONIAN INVARIANCE 185<br />

7.5 Example: L<strong>in</strong>ear harmonic oscillator on a cart mov<strong>in</strong>g at constant velocity<br />

Consider a l<strong>in</strong>ear harmonic oscillator located on a cart that<br />

is mov<strong>in</strong>g with constant velocity 0 <strong>in</strong> the direction, as shown<br />

<strong>in</strong> the adjacent figure. Let the laboratory frame be the unprimed<br />

frame, and the cart frame be designated the primed frame. Assume<br />

that = 0 at =0 Then<br />

0 = − 0 ˙ 0 = ˙ − 0 ¨ 0 =¨<br />

The harmonic oscillator will have a potential energy of<br />

= 1 2 02 = 1 2 ( − 0) 2<br />

x<br />

x’<br />

m<br />

Laboratory frame: The Lagrangian is<br />

v 0<br />

( ˙ ) = ˙2 − 1 v t<br />

0<br />

2 2 ( − 0) 2<br />

Lagrange equation Λ =0gives the equation of motion to be Harmonic oscillator on cart mov<strong>in</strong>g at<br />

¨ = −( − <br />

uniform velocity 0 .<br />

0 )<br />

The def<strong>in</strong>ition of generalized momentum gives<br />

= <br />

˙ = ˙<br />

The Hamiltonian is<br />

( ) = X <br />

˙ − = 2<br />

˙<br />

2 + 1 2 ( − 0) 2<br />

The Hamiltonian is the sum of the k<strong>in</strong>etic and potential energies and equals the total energy of the system,<br />

but it is not conserved s<strong>in</strong>ce and are both explicit functions of time, that is <br />

<br />

= <br />

<br />

= − <br />

<br />

6= 0.<br />

Physically this is understood <strong>in</strong> that energy must flow <strong>in</strong>to and out of the external constra<strong>in</strong>t keep<strong>in</strong>g the cart<br />

mov<strong>in</strong>g uniformly at a constant velocity 0 aga<strong>in</strong>st the reaction to the oscillat<strong>in</strong>g mass. That is, assum<strong>in</strong>g<br />

a uniform velocity for the mov<strong>in</strong>g cart constitutes a time-dependent constra<strong>in</strong>t on the mass, and the force of<br />

constra<strong>in</strong>t does work <strong>in</strong> actual displacement of the complete system. If the constra<strong>in</strong>t did not exist, then the<br />

cart momentum would oscillate such that the total momentum of cart plus spr<strong>in</strong>g system is conserved.<br />

Cart frame: Transform the Lagrangian to the primed coord<strong>in</strong>ates <strong>in</strong> the mov<strong>in</strong>g frame of reference,<br />

which also is an <strong>in</strong>ertial frame. Then the Lagrangian <strong>in</strong> terms of the mov<strong>in</strong>g cart frame coord<strong>in</strong>ates, is<br />

( 0 ˙ 0 )= ¡<br />

˙ 02 +2˙ 0 0 + 2 1<br />

2<br />

0¢<br />

−<br />

2 02<br />

The Lagrange equation of motion Λ 0 =0gives the equation of motion to be<br />

¨ 0 = − 0<br />

where 0 is the displacement of the mass with respect to the cart. This implies that an observer on the<br />

cart will observe simple harmonic motion as is to be expected from the pr<strong>in</strong>ciple of equivalence <strong>in</strong> Galilean<br />

relativity.<br />

The def<strong>in</strong>ition of the generalized momentum gives the l<strong>in</strong>ear momentum <strong>in</strong> the primed frame coord<strong>in</strong>ates<br />

to be<br />

0 = <br />

˙ 0 = ˙0 + 0<br />

The cart-frame Hamiltonian also can be expressed <strong>in</strong> terms of the coord<strong>in</strong>ates <strong>in</strong> the mov<strong>in</strong>g frame to be<br />

( 0 0 )= ˙ 0 <br />

˙ 0 − = (0 − 0 ) 2<br />

+ 1 2 2 02 − 2 2 0<br />

Note that the Lagrangian and Hamiltonian expressed <strong>in</strong> terms of the coord<strong>in</strong>ates <strong>in</strong> the cart frame of reference<br />

are not explicitly time dependent, therefore is conserved. However, the cart-frame Hamiltonian does not<br />

equal the total energy s<strong>in</strong>ce the coord<strong>in</strong>ate transformation is time dependent. Actually the first two terms <strong>in</strong><br />

the above Hamiltonian are the energy of the harmonic oscillator <strong>in</strong> the cart frame. This example shows that<br />

the Hamiltonians differ when expressed <strong>in</strong> terms of either the laboratory or cart frames of reference


186 CHAPTER 7. SYMMETRIES, INVARIANCE AND THE HAMILTONIAN<br />

7.6 Example: Isotropic central force <strong>in</strong> a rotat<strong>in</strong>g frame<br />

Consider a mass subject to a central isotropic radial<br />

force () as shown <strong>in</strong> the adjacent figure. Compare<br />

the Hamiltonian <strong>in</strong> the fixed frame of reference ,<br />

with the Hamiltonian 0 <strong>in</strong> a frame of reference 0 that<br />

is rotat<strong>in</strong>g about the center of the force with constant<br />

angular velocity Restrict this case to rotation about<br />

one axis so that only two polar coord<strong>in</strong>ates and need<br />

to be considered. The transformations are<br />

0 = <br />

0 = − <br />

z<br />

y<br />

Also<br />

() =( 0 )<br />

Fixed frame of reference :<br />

= − = 2<br />

x<br />

Mass subject to radial force<br />

³<br />

˙ 2 + ˙2´ 2 − ()<br />

m<br />

S<strong>in</strong>ce the Lagrangian is not explicitly time dependent, then the Hamiltonian is conserved. For this fixed-frame<br />

Hamiltonian the generalized momenta are<br />

= <br />

˙ = ˙2 ˙<br />

= <br />

˙ = ˙<br />

The Hamiltonian equals<br />

( )= X <br />

<br />

˙ − = 1 µ<br />

2 + <br />

2<br />

˙ 2 2 + () =<br />

The Hamiltonian <strong>in</strong> the fixed frame is conserved and equals the total energy, that is = + .<br />

Rotat<strong>in</strong>g frame of reference 0<br />

The above <strong>in</strong>ertial fixed-frame Lagrangian can be written <strong>in</strong> terms of the primed (non-<strong>in</strong>ertial rotat<strong>in</strong>g<br />

frame) coord<strong>in</strong>ates as<br />

µ<br />

˙ 02 + 02 ³ ˙0 + ´2<br />

− ( 0 )<br />

= − = ³<br />

˙ 2 + ˙2´<br />

2 − () = 2<br />

2<br />

The generalized momenta derived from this Lagrangian are<br />

0 = ³ ˙ 0 = ˙02 ˙0 + ´<br />

= 0 + 0 02 <br />

0 = <br />

˙ 0 = ˙02 = <br />

The Hamiltonian expressed <strong>in</strong> terms of the non-<strong>in</strong>ertial rotat<strong>in</strong>g frame coord<strong>in</strong>ates is<br />

⎛ ³ ´<br />

0 ( 0 0 0 0 )= <br />

˙ 0 ˙0 + <br />

˙ ˙ 0 0 − = 1 0<br />

⎝ 02 + ⎞<br />

+<br />

0 2 <br />

⎠<br />

2<br />

2 + ( 0 )<br />

Note that 0 ( 0 0 0 0 ) is time <strong>in</strong>dependent and therefore is conserved, but ( 0 0 0 0 ) 6= because<br />

the generalized coord<strong>in</strong>ates are time dependent. In addition, 0 <br />

is conserved s<strong>in</strong>ce<br />

0<br />

˙ 0 = <br />

0 = − <br />

0 =0


7.10. HAMILTONIAN INVARIANCE 187<br />

7.7 Example: The plane pendulum<br />

The simple plane pendulum <strong>in</strong> a uniform gravitational<br />

field is an example that illustrates Hamiltonian<br />

<strong>in</strong>variance. There is only one generalized coord<strong>in</strong>ate, <br />

and the Lagrangian for this system is<br />

= 1 2 2 ˙2 + cos <br />

The momentum conjugate to is<br />

g<br />

= <br />

˙ = 2 ˙<br />

which is the angular momentum about the pivot po<strong>in</strong>t.<br />

Us<strong>in</strong>g the Lagrange-Euler equation this gives that<br />

The plane pendulum constra<strong>in</strong>ed to oscillate <strong>in</strong> a<br />

vertical plane <strong>in</strong> a uniform gravitational field.<br />

<br />

= ˙ = = − s<strong>in</strong> <br />

<br />

Note that the angular momentum is not a constant of motion s<strong>in</strong>ce it explicitly depends on .<br />

The Hamiltonian is<br />

= X 1<br />

<br />

˙ − = ˙ 2 − =<br />

2 2 ˙2 − cos =<br />

<br />

− cos <br />

22 <br />

Note that the Lagrangian and Hamiltonian are not explicit functions of time, therefore they are conserved.<br />

Also the potential is velocity <strong>in</strong>dependent and there is no coord<strong>in</strong>ate transformation, thus the Hamiltonian<br />

equals the total energy which is a constant of motion.<br />

=<br />

2 <br />

− cos = <br />

22 m<br />

7.8 Example: Oscillat<strong>in</strong>g cyl<strong>in</strong>der <strong>in</strong> a cyl<strong>in</strong>drical bowl<br />

It is important to correctly account for constra<strong>in</strong>t forces when us<strong>in</strong>g<br />

Noether’s theorem for constra<strong>in</strong>ed systems. Noether’s theorem assumes<br />

the variables are <strong>in</strong>dependent. This is illustrated by consider<strong>in</strong>g<br />

the example of a solid cyl<strong>in</strong>der roll<strong>in</strong>g <strong>in</strong> a fixed cyl<strong>in</strong>drical bowl. Assume<br />

that a uniform cyl<strong>in</strong>der of radius and mass is constra<strong>in</strong>ed<br />

to roll without slipp<strong>in</strong>g on the <strong>in</strong>ner surface of the lower half of a hollow<br />

cyl<strong>in</strong>der of radius . The motion is constra<strong>in</strong>ed to ensure that<br />

the axes of both cyl<strong>in</strong>ders rema<strong>in</strong> parallel and .<br />

The generalized coord<strong>in</strong>ates are taken to be the angles and <br />

which are measured with respect to a fixed vertical axis. Then the<br />

k<strong>in</strong>etic energy and potential energy are<br />

= 1 h(<br />

2 − ) ˙<br />

i 2 1 +<br />

2 ˙ 2 =[ − ( − )cos] <br />

where is the mass of the small cyl<strong>in</strong>der and where =0at the lowest position of the sphere. The moment<br />

of <strong>in</strong>ertia of a uniform cyl<strong>in</strong>der is = 1 2 2 .<br />

The Lagrangian is<br />

− − = 1 h(<br />

2 − ) ˙<br />

i 2 1 +<br />

4 2 ˙2 − [ − ( − )cos] <br />

S<strong>in</strong>ce the solid cyl<strong>in</strong>der rotates without slipp<strong>in</strong>g <strong>in</strong>side the cyl<strong>in</strong>drical shell, then the equation of constra<strong>in</strong>t is<br />

() = − ( + ) =0


188 CHAPTER 7. SYMMETRIES, INVARIANCE AND THE HAMILTONIAN<br />

Us<strong>in</strong>g the Lagrangian, plus the one equation of constra<strong>in</strong>t, requires one Lagrange multiplier. Then the<br />

Lagrange equations of motion for and are<br />

<br />

− ∙ ¸ <br />

˙<br />

+ = 0<br />

<br />

<br />

− ∙ ¸ <br />

˙<br />

+ <br />

= 0<br />

Substitute the Lagrangian and the equation of constra<strong>in</strong>t gives two equations of motion<br />

− ( − ) s<strong>in</strong> − ( − ) 2 ¨ + ( − ) = 0<br />

The lower equation of motion gives that<br />

= − 1 2 ¨<br />

Substitute this <strong>in</strong>to the equation of constra<strong>in</strong>t gives<br />

− 1 2 2¨ − = 0<br />

= − 1 ( − ) ¨<br />

2<br />

Substitute this <strong>in</strong>to the first equation of motion gives the equation of motion for to be<br />

2<br />

¨ =<br />

3( − ) s<strong>in</strong> <br />

that is<br />

= − <br />

3 s<strong>in</strong> <br />

The torque act<strong>in</strong>g on the small cyl<strong>in</strong>der due to the frictional force is<br />

= 1 2 2¨ = −<br />

Thus the frictional force is<br />

= − = <br />

3 s<strong>in</strong> <br />

Noether’s theorem can be used to ascerta<strong>in</strong> if the angular momentum is a constant of motion. The<br />

derivative of the Lagrangian<br />

<br />

=( − ) s<strong>in</strong> <br />

<br />

and thus the Lagrange equations tells us that ˙ =( − ) s<strong>in</strong> . Therefore is not a constant of motion.<br />

The Lagrangian is not an explicit function of which would suggest that is a constant of motion.<br />

But this is <strong>in</strong>correct because the constra<strong>in</strong>t equation = (−)<br />

<br />

couples and , that is, they are not<br />

<strong>in</strong>dependent variables, and thus and are coupled by the constra<strong>in</strong>t equation. As a result is not a<br />

constant of motion because it is directly coupled to =( − ) s<strong>in</strong> which is not a constant of motion.<br />

Thus neither nor are constants of motion. This illustrates that one must account carefully for equations<br />

of constra<strong>in</strong>t, and the concomitant constra<strong>in</strong>t forces, when apply<strong>in</strong>g Noether’s theorem which tacitly assumes<br />

<strong>in</strong>dependent variables.<br />

The Hamiltonian can be derived us<strong>in</strong>g the generalized momenta<br />

= <br />

˙ = ( − )2 ˙<br />

= <br />

˙ = 1 2 2 ˙<br />

Then the Hamiltonian is given by<br />

2 <br />

= ˙ + ˙ − =<br />

2 ( − ) 2 + 2 <br />

+[ − ( − )cos] <br />

2 Note that the transformation to generalized coord<strong>in</strong>ates is time <strong>in</strong>dependent and the potential is not velocity<br />

dependent, thus the Hamiltonian also equals the total energy. Also the Hamiltonian is conserved s<strong>in</strong>ce<br />

<br />

=0.


7.11. HAMILTONIAN FOR CYCLIC COORDINATES 189<br />

7.11 Hamiltonian for cyclic coord<strong>in</strong>ates<br />

It is <strong>in</strong>terest<strong>in</strong>g to discuss the properties of the Hamiltonian for cyclic coord<strong>in</strong>ates for which <br />

<br />

=0.<br />

Ignor<strong>in</strong>g the external and Lagrange multiplier terms,<br />

˙ = = − =0 (7.49)<br />

<br />

That is, a cyclic coord<strong>in</strong>ate has a constant correspond<strong>in</strong>g momentum for the Hamiltonian as well as<br />

for the Lagrangian. Conversely, if a generalized coord<strong>in</strong>ate does not occur <strong>in</strong> the Hamiltonian, then the<br />

correspond<strong>in</strong>g generalized momentum is conserved. Cyclic coord<strong>in</strong>ates were discussed earlier when discuss<strong>in</strong>g<br />

symmetries and conservation-law aspects of the Lagrangian. For example, if the Lagrangian, or Hamiltonian<br />

do not depend on a l<strong>in</strong>ear coord<strong>in</strong>ate then is conserved. Similarly for and An extension of this<br />

pr<strong>in</strong>ciple has been derived for the relationship between time <strong>in</strong>dependence and total energy of a system,<br />

that is, the Hamiltonian equals the total energy if the transformation to generalized coord<strong>in</strong>ates is time<br />

<strong>in</strong>dependent and the potential is velocity <strong>in</strong>dependent.<br />

A valuable feature of the Hamiltonian formulation is that it allows elim<strong>in</strong>ation of cyclic variables which<br />

reduces the number of degrees of freedom to be handled. As a consequence, cyclic variables are called<br />

ignorable variables <strong>in</strong> Hamiltonian mechanics. For example, consider that the Lagrangian has one cyclic<br />

variable . As a consequence, the Lagrangian does not depend on , and thus it can be written as<br />

= ( 1 −1 ; ˙ 1 ˙ ; ) The Lagrangian still conta<strong>in</strong>s generalized velocities, thus one still has to<br />

treat degrees of freedom even though one degree of freedom is cyclic. However, <strong>in</strong> the Hamiltonian<br />

formulation, only − 1 degrees of freedom are required s<strong>in</strong>ce the momentum for the cyclic degree of freedom<br />

is a constant = Thus the Hamiltonian can be written as = ( 1 −1 ; 1 −1 ; ; ) ,thatis,<br />

the Hamiltonian <strong>in</strong>cludes only −1 degrees of freedom. Thus the dimension of the problem has been reduced<br />

by one s<strong>in</strong>ce the conjugate cyclic (ignorable) variables ( ) are elim<strong>in</strong>ated. Hamiltonian mechanics can<br />

significantly reduce the dimension of the problem when the system <strong>in</strong>volves several cyclic variables. This is<br />

<strong>in</strong> contrast to the situation for the Lagrangian approach as discussed <strong>in</strong> chapters 8 and 15.<br />

7.12 Symmetries and <strong>in</strong>variance<br />

This chapter has shown that the symmetries of a system lead to <strong>in</strong>variance of physical quantities as was proposed<br />

by Noether. The symmetry properties of the Lagrangian can lead to the conservation laws summarized<br />

<strong>in</strong> table 72.<br />

Table 7.2: Symmetries and conservation laws <strong>in</strong> classical mechanics<br />

Symmetry Lagrange property Conserved quantity<br />

Spatial <strong>in</strong>variance Translational <strong>in</strong>variance L<strong>in</strong>ear momentum<br />

Spatial homogeneous Rotational <strong>in</strong>variance Angular momentum<br />

Time <strong>in</strong>variance Time <strong>in</strong>dependence Total energy<br />

The importance of the relations between <strong>in</strong>variance and symmetry cannot be overemphasized. It extends<br />

beyond classical mechanics to quantum physics and field theory. For a three-dimensional closed system,<br />

there are three possible constants for l<strong>in</strong>ear momentum, three for angular momentum, and one for energy. It<br />

is especially <strong>in</strong>terest<strong>in</strong>g <strong>in</strong> that these, and only these, seven <strong>in</strong>tegrals have the property that they are additive<br />

for the particles compris<strong>in</strong>g a system, and this occurs <strong>in</strong>dependent of whether there is an <strong>in</strong>teraction among<br />

the particles. That is, this behavior is obeyed by the whole assemble of particles for f<strong>in</strong>ite systems. Because<br />

of its profound importance to physics, these relations between symmetry and <strong>in</strong>variance are used extensively.<br />

7.13 Hamiltonian <strong>in</strong> classical mechanics<br />

The Hamiltonian was def<strong>in</strong>ed by equation 737 dur<strong>in</strong>g the discussion of time <strong>in</strong>variance and energy conservation.<br />

The Hamiltonian is of much more profound importance to physics than implied by the ad hoc def<strong>in</strong>ition<br />

given by equation 737 This relates to the fact that the Hamiltonian is written <strong>in</strong> terms of the fundamental<br />

coord<strong>in</strong>ate and its generalized momentum def<strong>in</strong>ed by equation 73.


190 CHAPTER 7. SYMMETRIES, INVARIANCE AND THE HAMILTONIAN<br />

It is more convenient to write the generalized coord<strong>in</strong>ates plus their generalized momentum as<br />

vectors, e.g. q ≡ ( 1 2 ), p ≡ ( 1 2 ). The generalized momenta conjugate to the coord<strong>in</strong>ate ,<br />

def<strong>in</strong>ed by 73, thencanbewritten<strong>in</strong>theform<br />

(q ˙q t)<br />

= (7.50)<br />

˙ <br />

Substitut<strong>in</strong>g this def<strong>in</strong>ition of the generalized momentum <strong>in</strong>to the Hamiltonian def<strong>in</strong>ed <strong>in</strong> (737), and<br />

express<strong>in</strong>g it <strong>in</strong> terms of the coord<strong>in</strong>ate q and its conjugate generalized momenta p, leadsto<br />

(q p) = X <br />

˙ − (q ˙q) (7.51)<br />

= p · ˙q−(q ˙q) (7.52)<br />

Note that the scalar product p · ˙q = P ˙ equals 2 for systems that are scleronomic and when the<br />

potential is velocity <strong>in</strong>dependent.<br />

The crucial feature of the Hamiltonian is that it is expressed as (q p) that is, it is a function<br />

of the generalized coord<strong>in</strong>ates q and their conjugate momenta p, whicharetakentobe<strong>in</strong>dependent,<strong>in</strong><br />

addition to the <strong>in</strong>dependent variable, . This is <strong>in</strong> contrast to the Lagrangian (q ˙q) which is a function<br />

of the generalized coord<strong>in</strong>ates , the correspond<strong>in</strong>g velocities ˙ ,andtime The velocities ˙q are the<br />

time derivatives of the coord<strong>in</strong>ates q and thus these are related. In physics, the fundamental conjugate<br />

coord<strong>in</strong>ates are (q p) which are the coord<strong>in</strong>ates underly<strong>in</strong>g the Hamiltonian. This is <strong>in</strong> contrast to (q ˙q)<br />

which are the coord<strong>in</strong>ates that underlie the Lagrangian. Thus the Hamiltonian is more fundamental than<br />

the Lagrangian and is a reason why the Hamiltonian mechanics, rather than the Lagrangian mechanics, was<br />

used as the foundation for development of quantum and statistical mechanics.<br />

Hamiltonian mechanics will be derived two other ways. Chapter 8 uses the Legendre transformation<br />

between the conjugate variables (q ˙q) and (q p) where the generalized coord<strong>in</strong>ate q and its conjugate<br />

generalized momentum, p are <strong>in</strong>dependent. This shows that Hamiltonian mechanics is based on the same<br />

variational pr<strong>in</strong>ciples as those used to derive Lagrangian mechanics. Chapter 9 derives Hamiltonian mechanics<br />

directly from Hamilton’s Pr<strong>in</strong>ciple of Least action. Chapter 8 will <strong>in</strong>troduce the algebraic Hamiltonian<br />

mechanics, that is based on the Hamiltonian. The powerful capabilities provided by Hamiltonian mechanics<br />

will be described <strong>in</strong> chapter 15.<br />

7.14 Summary<br />

This chapter has explored the importance of symmetries and <strong>in</strong>variance <strong>in</strong> Lagrangian mechanics and has<br />

<strong>in</strong>troduced the Hamiltonian. The follow<strong>in</strong>g summarizes the important conclusions derived <strong>in</strong> this chapter.<br />

Noether’s theorem:<br />

Noether’s theorem explores the remarkable connection between symmetry, plus the <strong>in</strong>variance of a system<br />

under transformation, and related conservation laws which imply the existence of important physical<br />

pr<strong>in</strong>ciples, and constants of motion. Transformations where the equations of motion are <strong>in</strong>variant are called<br />

<strong>in</strong>variant transformations. Variables that are <strong>in</strong>variant to a transformation are called cyclic variables. It<br />

was shown that if the Lagrangian does not explicitly conta<strong>in</strong> a particular coord<strong>in</strong>ate of displacement, then<br />

the correspond<strong>in</strong>g conjugate momentum, ˙ is conserved. This is Noether’s theorem which states “For each<br />

symmetry of the Lagrangian, there is a conserved quantity”. In particular it was shown that translational<br />

<strong>in</strong>variance <strong>in</strong> a given direction leads to the conservation of l<strong>in</strong>ear momentum <strong>in</strong> that direction, and rotational<br />

<strong>in</strong>variance about an axis leads to conservation of angular momentum about that axis. These are the firstorder<br />

spatial and angular <strong>in</strong>tegrals of the equations of motion. Noether’s theorem also relates the properties<br />

of the Hamiltonian to time <strong>in</strong>variance of the Lagrangian, namely;<br />

(1) is conserved if, and only if, the Lagrangian, and consequently the Hamiltonian, are not explicit<br />

functions of time.<br />

(2) The Hamiltonian gives the total energy if the constra<strong>in</strong>ts and coord<strong>in</strong>ate transformations are time<br />

<strong>in</strong>dependent and the potential energy is velocity <strong>in</strong>dependent. This is equivalent to stat<strong>in</strong>g that = if the<br />

constra<strong>in</strong>ts, or generalized coord<strong>in</strong>ates, for the system are time <strong>in</strong>dependent.<br />

Noether’s theorem is of importance s<strong>in</strong>ce it underlies the relation between symmetries, and <strong>in</strong>variance <strong>in</strong><br />

all of physics; that is, its applicability extends beyond classical mechanics.


7.14. SUMMARY 191<br />

Generalized momentum:<br />

The generalized momentum associated with the coord<strong>in</strong>ate is def<strong>in</strong>ed to be<br />

<br />

≡ <br />

(73)<br />

˙ <br />

where is also called the conjugate momentum (or canonical momentum) to where are<br />

conjugate, or canonical, variables. Remember that the l<strong>in</strong>ear momentum is the first-order time <strong>in</strong>tegral<br />

given by equation 210. Note that if is not a spatial coord<strong>in</strong>ate, then is not l<strong>in</strong>ear momentum, but is<br />

the conjugate momentum. For example, if is an angle, then will be angular momentum.<br />

K<strong>in</strong>etic energy <strong>in</strong> generalized coord<strong>in</strong>ates:<br />

It was shown that the k<strong>in</strong>etic energy can be expressed <strong>in</strong> terms of generalized coord<strong>in</strong>ates by<br />

(q ˙q) = X X 1<br />

2 <br />

<br />

˙ ˙ + X X <br />

˙ + X X<br />

µ 2<br />

1<br />

<br />

<br />

<br />

<br />

<br />

<br />

2 <br />

(719)<br />

<br />

<br />

= 2 (q ˙q)+ 1 (q ˙q)+ 0 (q) (7.53)<br />

For scleronomic systems with a potential that is velocity <strong>in</strong>dependent, then the k<strong>in</strong>etic energy can be<br />

expressed as<br />

= 2 = 1 X<br />

˙ = 1 2 2 ˙q · p<br />

(731)<br />

Generalized energy<br />

Jacobi’s Generalized Energy (q ˙ ) was def<strong>in</strong>ed as<br />

(q ˙q) ≡ X µ <br />

<br />

˙ − (q ˙q)<br />

˙<br />

<br />

<br />

(736)<br />

Hamiltonian function<br />

The Hamiltonian (q p) was def<strong>in</strong>ed <strong>in</strong> terms of the generalized energy (q ˙q) and by <strong>in</strong>troduc<strong>in</strong>g<br />

the generalized momentum. That is<br />

(q p) ≡ (q ˙q)= X <br />

˙ − (q ˙q)=p · ˙q−(q ˙q)<br />

(737)<br />

Generalized energy theorem<br />

The equations of motion lead to the generalized energy theorem which states that the time dependence<br />

of the Hamiltonian is related to the time dependence of the Lagrangian.<br />

(q p)<br />

<br />

= X <br />

"<br />

#<br />

X<br />

˙ <br />

+ (q)<br />

<br />

=1<br />

−<br />

(q ˙q)<br />

<br />

(738)<br />

Note that if all the generalized non-potential forces are zero, then the bracket <strong>in</strong> equation 738 is zero, and<br />

if the Lagrangian is not an explicit function of time, then the Hamiltonian is a constant of motion.<br />

Generalized energy and total energy:<br />

The generalized energy, and correspond<strong>in</strong>g Hamiltonian, equal the total energy if:<br />

1) The k<strong>in</strong>etic energy has a homogeneous quadratic dependence on the generalized velocities and the<br />

transformation to generalized coord<strong>in</strong>ates is <strong>in</strong>dependent of time, <br />

<br />

=0<br />

2) The potential energy is not velocity dependent, thus the terms <br />

˙ <br />

=0<br />

Chapter 8 will <strong>in</strong>troduce Hamiltonian mechanics that is built on the Hamiltonian, and chapter 15 will<br />

explore applications of Hamiltonian mechanics.


192 CHAPTER 7. SYMMETRIES, INVARIANCE AND THE HAMILTONIAN<br />

Problems<br />

1. Consider a particle of mass mov<strong>in</strong>g <strong>in</strong> a plane and subject to an <strong>in</strong>verse square attractive force.<br />

(a) Obta<strong>in</strong> the equations of motion.<br />

(b) Is the angular momentum about the orig<strong>in</strong> conserved?<br />

(c) Obta<strong>in</strong> expressions for the generalized forces.<br />

2. Consider a Lagrangian function of the form ( ˙<br />

¨ ). Here the Lagrangian conta<strong>in</strong>s a time derivative<br />

of the generalized coord<strong>in</strong>ates that is higher than the first. When work<strong>in</strong>g with such Lagrangians, the term<br />

“generalized mechanics” is used.<br />

(a) Consider a system with one degree of freedom. By apply<strong>in</strong>g the methods of the calculus of variations,<br />

and assum<strong>in</strong>g that Hamilton’s pr<strong>in</strong>ciple holds with respect to variations which keep both and ˙ fixed at<br />

the end po<strong>in</strong>ts, show that the correspond<strong>in</strong>g Lagrange equation is<br />

2 µ <br />

2 ¨<br />

<br />

− <br />

µ <br />

˙<br />

<br />

+ <br />

=0<br />

Such equations of motion have <strong>in</strong>terest<strong>in</strong>g applications <strong>in</strong> chaos theory.<br />

(b) Apply this result to the Lagrangian<br />

Do you recognize the equations of motion?<br />

= − 2 ¨ − 2 2 <br />

3. A uniform solid cyl<strong>in</strong>der of radius and mass rests on a horizontal plane and an identical cyl<strong>in</strong>der rests<br />

on it touch<strong>in</strong>g along the top of the first cyl<strong>in</strong>der with the axes of both cyl<strong>in</strong>ders parallel. The upper cyl<strong>in</strong>der<br />

isgivenan<strong>in</strong>f<strong>in</strong>itessimal displacement so that both cyl<strong>in</strong>ders roll without slipp<strong>in</strong>g <strong>in</strong> the directions shown by<br />

the arrows.<br />

(a) F<strong>in</strong>d Lagrangian for this system<br />

(b) What are the constants of motion?<br />

(c) Show that as long as the cyl<strong>in</strong>ders rema<strong>in</strong> <strong>in</strong> contact then<br />

˙ 2 =<br />

12 (1 − cos )<br />

(17 + 4 cos − 4cos 2 )<br />

2<br />

1<br />

y<br />

x<br />

t = 0 t > 0<br />

4. Consider a diatomic molecule which has a symmetry axis along the l<strong>in</strong>e through the center of the two atoms<br />

compris<strong>in</strong>g the molecule. Consider that this molecule is rotat<strong>in</strong>g about an axis perpendicular to the symmetry<br />

axis and that there are no external forces act<strong>in</strong>g on the molecule. Use Noether’s Theorem to answer the<br />

follow<strong>in</strong>g questions:<br />

a) Is the total angular momentum conserved?<br />

b) Is the projection of the total angular momentum along a space-fixed axis conserved?<br />

c) Is the projection of the angular momentum along the symmetry axis of the rotat<strong>in</strong>g molecule conserved?<br />

d) Is the projection of the angular momentum perpendicular to the rotat<strong>in</strong>g symmetry axis conserved?


7.14. SUMMARY 193<br />

5. A bead of mass slides under gravity along a smooth wire bent <strong>in</strong> the shape of a parabola 2 = <strong>in</strong> the<br />

vertical ( ) plane.<br />

(a) What k<strong>in</strong>d (holonomic, nonholonomic, scleronomic, rheonomic) of constra<strong>in</strong>t acts on ?<br />

(b) Set up Lagrange’s equation of motion for with the constra<strong>in</strong>t embedded.<br />

(c) Set up Lagrange’s equations of motion for both and with the constra<strong>in</strong>t adjo<strong>in</strong>ed and a Lagrangian<br />

multiplier <strong>in</strong>troduced.<br />

(d) Show that the same equation of motion for results from either of the methods used <strong>in</strong> part (b) or part<br />

(c).<br />

(e) Express <strong>in</strong> terms of and ˙.<br />

(f) What are the and components of the force of constra<strong>in</strong>t <strong>in</strong> terms of and ˙?<br />

6. Let the horizontal plane be the − plane. A bead of mass is constra<strong>in</strong>ed to slide with speed along a<br />

curve described by the function = (). What force does the curve apply to the bead? (Ignore gravity)<br />

7. Consider the Atwoods mach<strong>in</strong>e shown. The masses are 4, 5, and3. Let and be the heights of the<br />

right two masses relative to their <strong>in</strong>itial positions.<br />

a) Solve this problem us<strong>in</strong>g the Euler-Lagrange equations<br />

b) Use Noether’s theorem to f<strong>in</strong>d the conserved momentum.<br />

4m<br />

x<br />

5m<br />

3m<br />

y<br />

8. Acubeofside2 and center of mass , isplacedonafixed horizontal cyl<strong>in</strong>der of radius and center as<br />

shown<strong>in</strong>thefigure. Orig<strong>in</strong>ally the cube is placed such that is centered above but it can roll from side to<br />

side without slipp<strong>in</strong>g. (a) Assum<strong>in</strong>g that use the Lagrangian approach to to f<strong>in</strong>d the frequency for small<br />

oscillations about the top of the cyl<strong>in</strong>der. For simplicity make the small angle approximation for before us<strong>in</strong>g<br />

the Lagrange-Euler equations. (b) What will be the motion if ? Note that the moment of <strong>in</strong>ertia of the<br />

cube about the center of mass is 2 3 2 .<br />

C<br />

b<br />

h<br />

O


194 CHAPTER 7. SYMMETRIES, INVARIANCE AND THE HAMILTONIAN<br />

9. Two equal masses of mass are glued to a massless hoop of radius is free to rotate about its center <strong>in</strong> a<br />

vertical plane. The angle between the masses is 2, as shown. F<strong>in</strong>d the frequency of oscillations.<br />

10. Three massless sticks each of length 2, andmass with the center of mass at the center of each stick, are<br />

h<strong>in</strong>ged at their ends as shown. The bottom end of the lower stick is h<strong>in</strong>ged at the ground. They are held so<br />

that the lower two sticks are vertical, and the upper one is tilted at a small angle with respect to the vertical.<br />

They are then released. At the <strong>in</strong>stant of release what are the three equations of motion derived from the<br />

Lagrangian derived assum<strong>in</strong>g that is small? Use these to determ<strong>in</strong>e the <strong>in</strong>itial angular accelerations of the<br />

three sticks.<br />

m<br />

m<br />

m


Chapter 8<br />

Hamiltonian mechanics<br />

8.1 Introduction<br />

The three major formulations of classical mechanics are<br />

1. Newtonian mechanics which is the most <strong>in</strong>tuitive vector formulation used <strong>in</strong> classical mechanics.<br />

2. Lagrangian mechanics is a powerful algebraic formulation of classical mechanics derived us<strong>in</strong>g either<br />

d’Alembert’s Pr<strong>in</strong>ciple, or Hamilton’s Pr<strong>in</strong>ciple. The latter states ”A dynamical system follows a path<br />

that m<strong>in</strong>imizes the time <strong>in</strong>tegral of the difference between the k<strong>in</strong>etic and potential energies”.<br />

3. Hamiltonian mechanics has a beautiful superstructure that, like Lagrangian mechanics, is built<br />

upon variational calculus, Hamilton’s pr<strong>in</strong>ciple, and Lagrangian mechanics.<br />

Hamiltonian mechanics is <strong>in</strong>troduced at this juncture s<strong>in</strong>ce it is closely <strong>in</strong>terwoven with Lagrange mechanics.<br />

Hamiltonian mechanics plays a fundamental role <strong>in</strong> modern physics, but the discussion of the important<br />

role it plays <strong>in</strong> modern physics will be deferred until chapters 15 and 18 where applications to modern physics<br />

are addressed.<br />

The follow<strong>in</strong>g important concepts were <strong>in</strong>troduced <strong>in</strong> chapter 7:<br />

The generalized momentum was def<strong>in</strong>ed to be given by<br />

(q ˙q)<br />

≡ (8.1)<br />

˙ <br />

Note that, as discussed <strong>in</strong> chapter 72, if the potential is velocity dependent, such as the Lorentz force, then<br />

the generalized momentum <strong>in</strong>cludes terms <strong>in</strong> addition to the usual mechanical momentum.<br />

Jacobi’s generalized energy function (q ˙q) was <strong>in</strong>troduced where<br />

X<br />

(q ˙q)=<br />

<br />

µ <br />

<br />

˙ − (q ˙q) (8.2)<br />

˙ <br />

The Hamiltonian function was def<strong>in</strong>ed to be given by express<strong>in</strong>g the generalized energy function,<br />

equation 82, <strong>in</strong> terms of the generalized momentum. That is, the Hamiltonian (q p) is expressed as<br />

X<br />

(q p)= ˙ − (q ˙q) (8.3)<br />

<br />

The symbols q, p, designate vectors of generalized coord<strong>in</strong>ates, q ≡ ( 1 2 ) p ≡ ( 1 2 ).<br />

Equation 83 can be written compactly <strong>in</strong> a symmetric form us<strong>in</strong>g the scalar product p · ˙q = P ˙ .<br />

(q p)+(q ˙q)=p · ˙q (8.4)<br />

A crucial feature of Hamiltonian mechanics is that the Hamiltonian is expressed as (q p) that<br />

is, it is a function of the generalized coord<strong>in</strong>ates and their conjugate momenta, which are taken to be<br />

<strong>in</strong>dependent, plus the <strong>in</strong>dependent variable, time. This contrasts with the Lagrangian (q ˙q) which is a<br />

function of the generalized coord<strong>in</strong>ates , and the correspond<strong>in</strong>g velocities ˙ , that is the time derivatives<br />

of the coord<strong>in</strong>ates , plus the <strong>in</strong>dependent variable, time.<br />

195


196 CHAPTER 8. HAMILTONIAN MECHANICS<br />

8.2 Legendre Transformation between Lagrangian and Hamiltonian<br />

mechanics<br />

Hamiltonian mechanics can be derived directly from Lagrange mechanics by consider<strong>in</strong>g the Legendre transformation<br />

between the conjugate variables (q ˙q) and (q p). Such a derivation is of considerable importance<br />

<strong>in</strong> that it shows that Hamiltonian mechanics is based on the same variational pr<strong>in</strong>ciples as those<br />

used to derive Lagrangian mechanics; that is d’Alembert’s Pr<strong>in</strong>ciple and Hamilton’s Pr<strong>in</strong>ciple. The general<br />

problem of convert<strong>in</strong>g Lagrange’s equations <strong>in</strong>to the Hamiltonian form h<strong>in</strong>ges on the <strong>in</strong>version of equation<br />

(81) that def<strong>in</strong>es the generalized momentum p This <strong>in</strong>version is simplified by the fact that (81) is the first<br />

partial derivative of the Lagrangian scalar function (q ˙q t).<br />

Asdescribed<strong>in</strong>appendix 4, consider transformations between two functions (u w) and (v w)<br />

where u and v are the active variables related by the functional form<br />

v = ∇ u (u w) (8.5)<br />

and where w designates passive variables. The function ∇ u (u w) is the first-order derivative, (gradient)<br />

of (u w) with respect to the components of the vector u. The Legendre transform states that the <strong>in</strong>verse<br />

formula can always be written as a first-order derivative<br />

The function (v w) is related to (u w) by the symmetric relation<br />

u = ∇ v (v w) (8.6)<br />

(v w)+ (u w) =u · v (8.7)<br />

where the scalar product u · v = P <br />

=1 .<br />

Furthermore the first-order derivatives with respect to all the passive variables are related by<br />

∇ w (u w) =−∇ w<br />

(v w) (8.8)<br />

The relationship between the functions (u w) and (v w) is symmetrical and each is said to be the<br />

Legendre transform of the other.<br />

The general Legendre transform can be used to relate the Lagrangian and Hamiltonian by identify<strong>in</strong>g the<br />

active variables v with p and u with ˙q the passive variable w with q, and the correspond<strong>in</strong>g functions<br />

(u w) =(q ˙q) and (v w) =(q p). Thus the generalized momentum (81) corresponds to<br />

p = ∇ ˙q (q ˙q) (8.9)<br />

where (q) are the passive variables. Then the Legendre transform states that the transformed variable ˙q<br />

is given by the relation<br />

˙q = ∇ p (q p) (8.10)<br />

S<strong>in</strong>ce the functions (q ˙q) and (q p) are the Legendre transforms of each other, they satisfy the<br />

relation<br />

(q p)+(q ˙q)=p · ˙q (8.11)<br />

The function (q p), which is the Legendre transform of the Lagrangian (q ˙q) is called the Hamiltonian<br />

function and equation (811) is identical to our orig<strong>in</strong>al def<strong>in</strong>ition of the Hamiltonian given by<br />

equation (83). Thevariablesq and are passive variables thus equation (88) gives that<br />

∇ q ( ˙q q) =−∇ q<br />

(p q) (8.12)<br />

Written <strong>in</strong> component form equation 812 gives the partial derivative relations<br />

( ˙q q) (p q)<br />

= −<br />

<br />

(8.13)<br />

( ˙q q) (p q)<br />

= −<br />

<br />

<br />

(8.14)


8.3. HAMILTON’S EQUATIONS OF MOTION 197<br />

Note that equations 813 and 814 are strictly a result of the Legendre transformation. To complete the<br />

transformation from Lagrangian to Hamiltonian mechanics it is necessary to <strong>in</strong>voke the calculus of variations<br />

via the Lagrange-Euler equations. The symmetry of the Legendre transform is illustrated by equation 811<br />

Equation 731 gives that the scalar product p · ˙q =2 2 For scleronomic systems, with velocity <strong>in</strong>dependent<br />

potentials the standard Lagrangian = − and =2 − + = + . Thus, for this simple<br />

case, equation 811 reduces to an identity + =2 .<br />

8.3 Hamilton’s equations of motion<br />

The explicit form of the Legendre transform 810 gives that the time derivative of the generalized coord<strong>in</strong>ate<br />

is<br />

(q p)<br />

˙ = (8.15)<br />

<br />

The Euler-Lagrange equation 660 is<br />

<br />

− X <br />

= + <br />

(8.16)<br />

˙ <br />

This gives the correspond<strong>in</strong>g Hamilton equation for the time derivative of to be<br />

=1<br />

<br />

= ˙ = X <br />

+ + <br />

(8.17)<br />

˙ <br />

Substitute equation 813 <strong>in</strong>to equation 817 leads to the second Hamilton equation of motion<br />

(q p)<br />

˙ = − +<br />

<br />

=1<br />

X<br />

=1<br />

<br />

<br />

<br />

+ <br />

(8.18)<br />

One can explore further the implications of Hamiltonian mechanics by tak<strong>in</strong>g the time differential of (83)<br />

giv<strong>in</strong>g.<br />

(q p)<br />

= X µ<br />

<br />

˙ <br />

<br />

+ ˙ <br />

<br />

− <br />

<br />

<br />

− <br />

˙ <br />

− <br />

(8.19)<br />

˙ <br />

Insert<strong>in</strong>g the conjugate momenta ≡ <br />

˙ <br />

Ã<br />

and equation 817 <strong>in</strong>to equation 819 results <strong>in</strong><br />

# !<br />

X <br />

− ˙ <br />

˙ − <br />

<br />

(q p)<br />

= X "<br />

˙ <br />

˙ ˙ + <br />

<br />

− ˙ −<br />

<br />

=1<br />

The second and fourth terms cancel as well as the ˙ ˙ terms, leav<strong>in</strong>g<br />

(q p)<br />

= X Ã" <br />

# !<br />

X <br />

+ <br />

˙ <br />

<br />

<br />

<br />

<br />

This is the generalized energy theorem given by equation 738.<br />

The total differential of the Hamiltonian also can be written as<br />

(q p)<br />

<br />

= X <br />

Use equations 815 and 818 to substitute for <br />

<br />

Ã"<br />

X <br />

(q p)<br />

<br />

= X <br />

=1<br />

=1<br />

µ <br />

˙ + <br />

˙ + <br />

<br />

and <br />

<br />

<br />

+ <br />

<br />

<br />

− <br />

<br />

<strong>in</strong> equation 822 gives<br />

# !<br />

˙ +<br />

(q p)<br />

<br />

− <br />

<br />

(8.20)<br />

(8.21)<br />

(8.22)<br />

(8.23)


198 CHAPTER 8. HAMILTONIAN MECHANICS<br />

Note that equation 823 must equal the generalized energy theorem, i.e. equation 821 Therefore,<br />

<br />

= − <br />

In summary, Hamilton’s equations of motion are given by<br />

(8.24)<br />

(q p)<br />

˙ = (8.25)<br />

<br />

" <br />

#<br />

(q p) X <br />

˙ = − + + <br />

<br />

(8.26)<br />

<br />

=1<br />

(q p)<br />

= X Ã" <br />

# !<br />

X <br />

+ (q ˙q)<br />

˙ − (8.27)<br />

<br />

<br />

<br />

<br />

=1<br />

The symmetry of Hamilton’s equations of motion is illustrated when the Lagrange multiplier and generalized<br />

forces are zero. Then<br />

(p q)<br />

<br />

(q p)<br />

˙ = (8.28)<br />

<br />

(p q)<br />

˙ = − (8.29)<br />

<br />

=<br />

(p q)<br />

<br />

( ˙q q)<br />

= −<br />

<br />

(8.30)<br />

This simplified form illustrates the symmetry of Hamilton’s equations of motion. Many books present<br />

the Hamiltonian only for this special simplified case where it is holonomic, conservative, and generalized<br />

coord<strong>in</strong>ates are used.<br />

8.3.1 Canonical equations of motion<br />

Hamilton’s equations of motion, summarized <strong>in</strong> equations 825 − 27 use either a m<strong>in</strong>imal set of generalized<br />

coord<strong>in</strong>ates, or the Lagrange multiplier terms, to account for holonomic constra<strong>in</strong>ts, or generalized forces<br />

<br />

to account for non-holonomic or other forces. Hamilton’s equations of motion usually are called the<br />

canonical equations of motion. Note that the term “canonical” has noth<strong>in</strong>g to do with religion or canon<br />

law; the reason for this name has bewildered many generations of students of classical mechanics. The<br />

term was <strong>in</strong>troduced by Jacobi <strong>in</strong> 1837 to designate a simple and fundamental set of conjugate variables<br />

and equations. Note the symmetry of Hamilton’s two canonical equations, plus the fact that the canonical<br />

variables are treated as <strong>in</strong>dependent canonical variables. The Lagrange mechanics coord<strong>in</strong>ates (q ˙q)<br />

are replaced by the Hamiltonian mechanics coord<strong>in</strong>ates (q p) where the conjugate momenta p are taken<br />

to be <strong>in</strong>dependent of the coord<strong>in</strong>ate q.<br />

Lagrange was the first to derive the canonical equations but he did not recognize them as a basic set of<br />

equations of motion. Hamilton derived the canonical equations of motion from his fundamental variational<br />

pr<strong>in</strong>ciple, chapter 92, and made them the basis for a far-reach<strong>in</strong>g theory of dynamics. Hamilton’s equations<br />

give 2 first-order differential equations for for each of the = − degrees of freedom. Lagrange’s<br />

equations give second-order differential equations for the <strong>in</strong>dependent generalized coord<strong>in</strong>ates ˙ <br />

It has been shown that (p q) and ( ˙q q) are the Legendre transforms of each other. Although<br />

the Lagrangian formulation is ideal for solv<strong>in</strong>g numerical problems <strong>in</strong> classical mechanics, the Hamiltonian<br />

formulation provides a better framework for conceptual extensions to other fields of physics s<strong>in</strong>ce it is written<br />

<strong>in</strong> terms of the fundamental conjugate coord<strong>in</strong>ates, q p. The Hamiltonian is used extensively <strong>in</strong> modern<br />

physics, <strong>in</strong>clud<strong>in</strong>g quantum physics, as discussed <strong>in</strong> chapters 15 and 18. For example, <strong>in</strong> quantum mechanics<br />

there is a straightforward relation between the classical and quantal representations of momenta; this does<br />

not exist for the velocities.<br />

The concept of state space, <strong>in</strong>troduced <strong>in</strong> chapter 332, applies naturally to Lagrangian mechanics s<strong>in</strong>ce<br />

( ˙ ) are the generalized coord<strong>in</strong>ates used <strong>in</strong> Lagrangian mechanics. The concept of Phase Space, <strong>in</strong>troduced<br />

<strong>in</strong> chapter 333, naturally applies to Hamiltonian phase space s<strong>in</strong>ce ( ) are the generalized coord<strong>in</strong>ates<br />

used <strong>in</strong> Hamiltonian mechanics.


8.4. HAMILTONIAN IN DIFFERENT COORDINATE SYSTEMS 199<br />

8.4 Hamiltonian <strong>in</strong> different coord<strong>in</strong>ate systems<br />

Prior to solv<strong>in</strong>g problems us<strong>in</strong>g Hamiltonian mechanics, it is useful to express the Hamiltonian <strong>in</strong> cyl<strong>in</strong>drical<br />

and spherical coord<strong>in</strong>ates for the special case of conservative forces s<strong>in</strong>ce these are encountered frequently<br />

<strong>in</strong> physics.<br />

8.4.1 Cyl<strong>in</strong>drical coord<strong>in</strong>ates <br />

Consider cyl<strong>in</strong>drical coord<strong>in</strong>ates Expressed <strong>in</strong> cartesian coord<strong>in</strong>ate<br />

= cos (8.31)<br />

= s<strong>in</strong> <br />

= <br />

Us<strong>in</strong>g appendix table 3 the Lagrangian can be written <strong>in</strong> cyl<strong>in</strong>drical coord<strong>in</strong>ates as<br />

= − = ³˙ 2 + 2 ˙2 + ˙<br />

2´<br />

− ( ) (8.32)<br />

2<br />

The conjugate momenta are<br />

= = ˙<br />

˙<br />

(8.33)<br />

= <br />

˙ = 2 ˙ (8.34)<br />

= = ˙<br />

˙<br />

(8.35)<br />

Assume a conservative force, then is conserved. S<strong>in</strong>ce the transformation from cartesian to nonrotat<strong>in</strong>g<br />

generalized cyl<strong>in</strong>drical coord<strong>in</strong>ates is time <strong>in</strong>dependent, then = Then us<strong>in</strong>g (832 − 835) gives<br />

the Hamiltonian <strong>in</strong> cyl<strong>in</strong>drical coord<strong>in</strong>ates to be<br />

(q p) = X ˙ − (q ˙q) (8.36)<br />

<br />

³<br />

´<br />

= ˙ + ˙ + ˙ − µ<br />

<br />

<br />

2 + 2 2 + 2 + ( )<br />

2<br />

Ã<br />

!<br />

= 1 2 + 2 <br />

2 2 + 2 + ( ) (8.37)<br />

The canonical equations of motion <strong>in</strong> cyl<strong>in</strong>drical coord<strong>in</strong>ates can be written as<br />

˙ = − <br />

=<br />

2 <br />

3 − <br />

<br />

(8.38)<br />

˙ = − <br />

= − <br />

(8.39)<br />

˙ = − <br />

= − <br />

(8.40)<br />

˙ = = <br />

<br />

(8.41)<br />

˙ = =<br />

<br />

2 (8.42)<br />

˙ = = <br />

<br />

(8.43)<br />

Note that if is cyclic, that is <br />

=0 then the angular momentum about the axis, ,isaconstant<br />

of motion. Similarly, if is cyclic, then is a constant of motion.


200 CHAPTER 8. HAMILTONIAN MECHANICS<br />

8.4.2 Spherical coord<strong>in</strong>ates, <br />

Appendix table 4 shows that the spherical coord<strong>in</strong>ates are related to the cartesian coord<strong>in</strong>ates by<br />

= s<strong>in</strong> cos (8.44)<br />

= s<strong>in</strong> s<strong>in</strong> <br />

= cos <br />

The Lagrangian is<br />

= − = 2<br />

³<br />

˙ 2 + 2 ˙2 + 2 s<strong>in</strong> 2 ˙ 2´ − () (8.45)<br />

The conjugate momenta are<br />

= <br />

= ˙ (8.46)<br />

= <br />

= 2 ˙ (8.47)<br />

= <br />

= 2 s<strong>in</strong> 2 ˙ (8.48)<br />

Assum<strong>in</strong>g a conservative force then is conserved. S<strong>in</strong>ce the transformation from cartesian to generalized<br />

spherical coord<strong>in</strong>ates is time <strong>in</strong>dependent, then = Thus us<strong>in</strong>g (846 − 848) the Hamiltonian is given<br />

<strong>in</strong> spherical coord<strong>in</strong>ates by<br />

(q p) = X ˙ − (q ˙q) (8.49)<br />

<br />

³<br />

´<br />

= ˙ + ˙ + ˙ − ³<br />

˙ 2 + 2 ˙2 + 2 s<strong>in</strong> 2 <br />

2<br />

˙ 2´ + ( ) (8.50)<br />

Ã<br />

!<br />

= 1 2 + 2 <br />

2 2 +<br />

2 <br />

2 s<strong>in</strong> 2 + ( ) (8.51)<br />

<br />

Then the canonical equations of motion <strong>in</strong> spherical coord<strong>in</strong>ates are<br />

à !<br />

˙ = − <br />

= 1<br />

3 2 +<br />

2 <br />

s<strong>in</strong> 2 − <br />

<br />

˙ = − <br />

= 1<br />

2 Ã<br />

<br />

2<br />

<br />

cos <br />

s<strong>in</strong> 3 <br />

!<br />

− <br />

<br />

(8.52)<br />

(8.53)<br />

˙ = − <br />

= − <br />

(8.54)<br />

˙ = = <br />

<br />

(8.55)<br />

˙ = =<br />

<br />

2 (8.56)<br />

˙ = <br />

=<br />

2 s<strong>in</strong> 2 <br />

(8.57)<br />

Note that if the coord<strong>in</strong>ate is cyclic, that is <br />

=0then the angular momentum is conserved. Also<br />

if the coord<strong>in</strong>ate is cyclic, and =0 that is, there is no change <strong>in</strong> the angular momentum perpendicular<br />

to the axis, then is conserved.<br />

An especially important spherically-symmetric Hamiltonian is that for a central field. Central fields, such<br />

as the gravitational or Coulomb fields of a uniform spherical mass, or charge, distributions, are spherically<br />

symmetric and then both and are cyclic. Thus the projection of the angular momentum about the <br />

axis is conserved for these spherically symmetric potentials. In addition, s<strong>in</strong>ce both and are conserved,<br />

then the total angular momentum also must be conserved as is predicted by Noether’s theorem


8.5. APPLICATIONS OF HAMILTONIAN DYNAMICS 201<br />

8.5 Applications of Hamiltonian Dynamics<br />

The equations of motion of a system can be derived us<strong>in</strong>g the Hamiltonian coupled with Hamilton’s equations<br />

of motion, that is, equations 825 − 827.<br />

Formally the Hamiltonian is constructed from the Lagrangian. That is<br />

1) Select a set of <strong>in</strong>dependent generalized coord<strong>in</strong>ates <br />

2) Partition the active forces.<br />

3) Construct the Lagrangian ( ˙ )<br />

4) Derive the conjugate generalized momenta via = <br />

˙ <br />

5) Know<strong>in</strong>g ˙ derive = P ˙ − <br />

6) Derive ˙ = <br />

<br />

and ˙ = − (qp)<br />

<br />

+ P <br />

=1 <br />

<br />

+ <br />

<br />

This procedure appears to be unnecessarily complicated compared to just us<strong>in</strong>g the Lagrangian plus<br />

Lagrangian mechanics to derive the equations of motion. Fortunately the above lengthy procedure often can<br />

be bypassed for conservative systems. That is, if the follow<strong>in</strong>g conditions are satisfied;<br />

) = ( ) − (), thatis, () is <strong>in</strong>dependent of the velocity ˙.<br />

) the generalized coord<strong>in</strong>ates are time <strong>in</strong>dependent.<br />

then it is possible to use the fact that = + = .<br />

The follow<strong>in</strong>g five examples illustrate the use of Hamiltonian mechanics to derive the equations of motion.<br />

8.1 Example: Motion <strong>in</strong> a uniform gravitational field<br />

Consider a mass <strong>in</strong> a uniform gravitational field act<strong>in</strong>g <strong>in</strong> the −z direction.<br />

simple case is<br />

= 1 2 ¡ ˙ 2 + ˙ 2 + ˙ 2¢ − <br />

The Lagrangian for this<br />

Therefore the generalized momenta are = <br />

˙ = ˙ = <br />

˙ = ˙ = <br />

˙<br />

Hamiltonian is<br />

= ˙. The correspond<strong>in</strong>g<br />

= X <br />

˙ − = ˙ + ˙ + ˙ − <br />

= 2 <br />

+ 2 <br />

+ 2 <br />

− 1 2<br />

Ã<br />

!<br />

Ã<br />

!<br />

2 <br />

+ 2 <br />

+ 2 <br />

+ = 1 2 <br />

<br />

2 + 2 <br />

+ 2 <br />

+ <br />

<br />

Note that the Lagrangian is not explicitly time dependent, thus the Hamiltonian is a constant of motion.<br />

Hamilton’s equations give that<br />

˙ = = <br />

<br />

˙ = = <br />

<br />

˙ = = <br />

<br />

− ˙ = <br />

=0<br />

− ˙ = <br />

=0<br />

− ˙ = <br />

= <br />

Comb<strong>in</strong><strong>in</strong>g these gives that ¨ =0 ¨ =0 ¨ = −. Note that the l<strong>in</strong>ear momenta and are constants<br />

of motion whereas the rate of change of is given by the gravitational force . Notealsothat = + <br />

for this conservative system.<br />

8.2 Example: One-dimensional harmonic oscillator<br />

Consider a mass subject to a l<strong>in</strong>ear restor<strong>in</strong>g force with spr<strong>in</strong>g constant The Lagrangian = − <br />

equals<br />

= 1 2 ˙2 − 1 2 2<br />

Therefore the generalized momentum is<br />

= <br />

˙ = ˙


202 CHAPTER 8. HAMILTONIAN MECHANICS<br />

The Hamiltonian is<br />

= X <br />

˙ − = ˙ − <br />

= <br />

− 1 2 <br />

2 + 1 2 2 = 1 2 <br />

2 + 1 2 2<br />

Note that the Lagrangian is not explicitly time dependent, thus the Hamiltonian will be a constant of motion.<br />

Hamilton’s equations give that<br />

˙ = = <br />

<br />

or<br />

= ˙<br />

In addition<br />

Comb<strong>in</strong><strong>in</strong>g these gives that<br />

− ˙ = <br />

= <br />

= <br />

¨ + =0<br />

which is the equation of motion for the harmonic oscillator.<br />

8.3 Example: Plane pendulum<br />

The plane pendulum, <strong>in</strong> a uniform gravitational field is an <strong>in</strong>terest<strong>in</strong>g system to consider. There is<br />

only one generalized coord<strong>in</strong>ate, and the Lagrangian for this system is<br />

The momentum conjugate to is<br />

= 1 2 2 ˙2 + cos <br />

= <br />

˙ = 2 ˙<br />

which is the angular momentum about the pivot po<strong>in</strong>t.<br />

The Hamiltonian is<br />

= X <br />

1<br />

<br />

˙ − = ˙ 2 − =<br />

2 2 ˙2 − cos =<br />

<br />

− cos <br />

22 Hamilton’s equations of motion give<br />

˙ = <br />

<br />

= <br />

2<br />

<br />

˙ = − = − s<strong>in</strong> <br />

<br />

Note that the Lagrangian and Hamiltonian are not explicit functions of time, therefore they are conserved.<br />

Also the potential is velocity <strong>in</strong>dependent and there is no coord<strong>in</strong>ate transformation, thus the Hamiltonian<br />

equals the total energy, that is<br />

=<br />

2 <br />

− cos = <br />

22 where is a constant of motion. Note that the angular momentum is not a constant of motion s<strong>in</strong>ce ˙ <br />

explicitly depends on .


8.5. APPLICATIONS OF HAMILTONIAN DYNAMICS 203<br />

The solutions for the plane pendulum on a ( ) phase diagram,<br />

shown <strong>in</strong> the adjacent figure, illustrate the motion. The<br />

upper phase-space plot shows the range ( = ± ). Note that<br />

the =+ and − correspond to the same physical po<strong>in</strong>t, that is<br />

the phase diagram should be rolled <strong>in</strong>to a cyl<strong>in</strong>der connected along<br />

the dashed l<strong>in</strong>es. The lower phase space plot shows two cycles for<br />

to better illustrate the cyclic nature of the phase diagram. The<br />

correspond<strong>in</strong>g state-space diagram is shown <strong>in</strong> figure 34. The<br />

trajectories are ellipses for low energy − correspond<strong>in</strong>g<br />

to oscillations of the pendulum about =0.Thecenter<br />

of the ellipse (0 0) is a stable equilibrium po<strong>in</strong>t for the oscillation.<br />

However,thereisaphasechangetorotationalmotionaboutthe<br />

horizontal axis when || ,thatis,thependulumsw<strong>in</strong>gs<br />

around a circle cont<strong>in</strong>uously, i.e. it rotates cont<strong>in</strong>uously <strong>in</strong> one<br />

direction about the horizontal axis. The phase change occurs at<br />

= and is designated by the separatrix trajectory.<br />

The plot of versus for the plane pendulum is better presented<br />

on a cyl<strong>in</strong>drical phase space representation s<strong>in</strong>ce is a<br />

cyclic variable that cycles around the cyl<strong>in</strong>der, whereas oscillates<br />

equally about zero hav<strong>in</strong>g both positive and negative values.<br />

When wrapped around a cyl<strong>in</strong>der then the unstable and stable<br />

equilibrium po<strong>in</strong>ts will be at diametrically opposite locations on<br />

the surface of the cyl<strong>in</strong>der at = 0. For small oscillations<br />

about equilibrium, also called librations, the correlation between<br />

and is given by the clockwise closed ellipses wrapped on the<br />

cyl<strong>in</strong>drical surface, whereas for energies || the positive<br />

corresponds to counterclockwise rotations while the negative<br />

corresponds to clockwise rotations.<br />

P<br />

O<br />

P<br />

(b)<br />

Elliptic po<strong>in</strong>t<br />

Hyperbolic po<strong>in</strong>t<br />

Phase-space diagrams for the plane<br />

pendulum. The separatrix (bold l<strong>in</strong>e)<br />

separates the oscillatory solutions from<br />

the roll<strong>in</strong>g solutions. The upper (a)<br />

shows one complete cycle while the lower<br />

(b) shows two complete cycles.<br />

8.4 Example: Hooke’s law force constra<strong>in</strong>ed to the surface of a cyl<strong>in</strong>der<br />

Consider the case where a mass is attracted by a<br />

force directed toward the orig<strong>in</strong> and proportional to the<br />

distance from the orig<strong>in</strong>. Determ<strong>in</strong>e the Hamiltonian<br />

if the mass is constra<strong>in</strong>ed to move on the surface of a<br />

cyl<strong>in</strong>der def<strong>in</strong>ed by<br />

2 + 2 = 2<br />

z<br />

It is natural to transform this problem to cyl<strong>in</strong>drical coord<strong>in</strong>ates<br />

. S<strong>in</strong>ce the force is just Hooke’s law<br />

F = −r<br />

the potential is the same as for the harmonic oscillator,<br />

that is<br />

= 1 2 2 = 1 2 (2 + 2 )<br />

This is <strong>in</strong>dependent of and thus is cyclic.<br />

In cyl<strong>in</strong>drical coord<strong>in</strong>ates the velocity is<br />

2 = ˙ 2 + 2 ˙2 +<br />

<br />

<br />

2<br />

x<br />

z<br />

Mass attracted to orig<strong>in</strong> by force proportional to<br />

distance from orig<strong>in</strong> with the motion constra<strong>in</strong>ed<br />

to the surface of a cyl<strong>in</strong>der.<br />

y<br />

Conf<strong>in</strong>ed to the surface of the cyl<strong>in</strong>der means that<br />

= <br />

<br />

= 0


204 CHAPTER 8. HAMILTONIAN MECHANICS<br />

Then the Lagrangian simplifies to<br />

= − = 1 2 ³<br />

2 ˙2 + ˙<br />

2´ − 1 2 (2 + 2 )<br />

The generalized coord<strong>in</strong>ates are and the correspond<strong>in</strong>g generalized momenta are<br />

= <br />

˙ = 2 ˙<br />

(a)<br />

= <br />

˙ = ˙<br />

(b)<br />

The system is conservative, and the transformation from rectangular to cyl<strong>in</strong>drical coord<strong>in</strong>ates does not<br />

depend explicitly on time. Therefore the Hamiltonian is conserved and equals the total energy. That is<br />

= X <br />

˙ − =<br />

2 <br />

2 2 + 2 <br />

2 + 1 2 (2 + 2 )=<br />

The equations of motion then are given by the canonical equations<br />

˙ = − <br />

=0<br />

˙ = − <br />

= −<br />

Equation(a)and(c)implythat<br />

<br />

˙ = =<br />

<br />

2<br />

<br />

˙ = = <br />

<br />

= <br />

= 2 ˙ = constant<br />

Thus the angular momentum about the axis of the cyl<strong>in</strong>der is conserved, that is, it is a cyclic variable.<br />

Comb<strong>in</strong><strong>in</strong>g equations (b) and (d) implies that<br />

(c)<br />

(d)<br />

¨ + =0<br />

q<br />

<br />

This is the equation for simple harmonic motion with angular frequency =<br />

<br />

. The symmetries imply<br />

that this problem has the same solutions for the coord<strong>in</strong>ate as the harmonic oscillator, while the coord<strong>in</strong>ate<br />

moves with constant angular velocity.<br />

8.5 Example: Electron motion <strong>in</strong> a cyl<strong>in</strong>drical magnetron<br />

A magnetron comprises a hot cyl<strong>in</strong>drical wire cathode that emits electrons and is at a high negative<br />

voltage. It is surrounded by a larger diameter concentric cyl<strong>in</strong>drical anode at ground potential. A uniform<br />

magnetic field runs parallel to the cyl<strong>in</strong>drical axis of the magnetron. The electron beam excites a multiple set<br />

of microwave cavities located around the circumference of the cyl<strong>in</strong>drical wall of the anode. The magnetron<br />

was <strong>in</strong>vented <strong>in</strong> England dur<strong>in</strong>g World War 2 to generate microwaves required for the development of radar.<br />

Consider a non-relativistic electron of mass and charge − <strong>in</strong> a cyl<strong>in</strong>drical magnetron mov<strong>in</strong>g between<br />

the central cathode wire, of radius at a negative electric potential − 0 , and a concentric cyl<strong>in</strong>drical anode<br />

conductor of radius which has zero electric potential. There is a uniform constant magnetic field parallel<br />

to the cyl<strong>in</strong>drical axis of the magnetron.<br />

Us<strong>in</strong>g SI units and cyl<strong>in</strong>drical coord<strong>in</strong>ates ( ) aligned with the axis of the magnetron, the electromagnetic<br />

force Lagrangian, given <strong>in</strong> chapter 610 equals<br />

= 1 2 ṙ2 + ( − ṙ · A)<br />

The electric and vector potentials for the magnetron geometry are<br />

ln( <br />

= − )<br />

0<br />

ln( )<br />

A = 1 2 ˆ


8.5. APPLICATIONS OF HAMILTONIAN DYNAMICS 205<br />

Thus expressed <strong>in</strong> cyl<strong>in</strong>drical coord<strong>in</strong>ates the Lagrangian equals<br />

= 1 2 ³<br />

˙ 2 + 2 ˙2 + ˙<br />

2´ + − 1 2 2 ˙<br />

The generalized momenta are<br />

= <br />

˙ = ˙<br />

= <br />

˙ = 1<br />

2 ˙ −<br />

2 2<br />

= <br />

˙ = ˙<br />

Note that the vector potential contributes an additional term to the angular momentum .<br />

Us<strong>in</strong>g the above generalized momenta leads to the Hamiltonian<br />

= ˙ + ˙ + ˙ − <br />

= 1 ³<br />

2 ˙ 2 + 2 ˙2 + ˙<br />

2´ − + 1 2 2 ˙<br />

= 2 <br />

2 + 1 µ<br />

2 2 + 1 2<br />

2 2 + 2 <br />

2 − <br />

= 1<br />

2<br />

"<br />

2 +<br />

µ<br />

<br />

+ 1 2 2<br />

+ 2 <br />

#<br />

− <br />

Note that the Hamiltonian is not an explicit function of time, therefore it is a constant of motion which<br />

equals the total energy.<br />

"<br />

= 1 µ<br />

2 <br />

+<br />

2 + 1 # 2<br />

2 + 2 − = <br />

S<strong>in</strong>ce ˙ = − <br />

<br />

and if isnotanexplicitfunctionof then ˙ =0 that is, is a constant of motion.<br />

Thus and are constants of motion.<br />

Consider the <strong>in</strong>itial conditions = ˙ = ˙ = ˙ =0.Then<br />

= <br />

˙ = 1<br />

2 ˙ −<br />

2 2 = − 1 2 2<br />

= 0<br />

= 1<br />

2<br />

"<br />

2 +<br />

µ<br />

<br />

+ 1 # 2<br />

2 + 2 ln( <br />

+ )<br />

0<br />

ln( ) = 0<br />

Note that at = then is given by the last equation s<strong>in</strong>ce the Hamiltonian equals a constant 0 .That<br />

is, assum<strong>in</strong>g that then<br />

2 =2 0 − ( 1 2 )2<br />

Def<strong>in</strong>e a critical magnetic field by<br />

then<br />

≡ 2 r<br />

20<br />

<br />

¡<br />

<br />

2<br />

<br />

¢= = ¡ 2 − 2¢ ( 1 2 )2<br />

Note that if then is real at = . However, if then is imag<strong>in</strong>ary at = <br />

imply<strong>in</strong>g that there must be a maximum orbit radius 0 for the electron where 0 . That is, the electron<br />

trajectories are conf<strong>in</strong>ed spatially to coaxial cyl<strong>in</strong>drical orbits concentric with the magnetron electromagnetic<br />

fields. These closed electron trajectories excite the microwave cavities located <strong>in</strong> the nearby outer cyl<strong>in</strong>drical<br />

wall of the anode.<br />

.


206 CHAPTER 8. HAMILTONIAN MECHANICS<br />

8.6 Routhian reduction<br />

Noether’s theorem states that if the coord<strong>in</strong>ate is cyclic, and if the Lagrange multiplier plus generalized<br />

force contributions for the coord<strong>in</strong>ates are zero, then the canonical momentum of the cyclic variable, is<br />

a constant of motion as is discussed <strong>in</strong> chapter 73. Therefore, both ( ) are constants of motion for cyclic<br />

variables, and these constant ( ) coord<strong>in</strong>ates can be factored out of the Hamiltonian (p q). This<br />

reduces the number of degrees of freedom <strong>in</strong>cluded <strong>in</strong> the Hamiltonian. For this reason, cyclic variables are<br />

called ignorable variables <strong>in</strong> Hamiltonian mechanics. This advantage does not apply to the ( ˙ ) variables<br />

used <strong>in</strong> Lagrangian mechanics s<strong>in</strong>ce ˙ is not a constant of motion for a cyclic coord<strong>in</strong>ate. The ability<br />

to elim<strong>in</strong>ate the cyclic variables as unknowns <strong>in</strong> the Hamiltonian is a valuable advantage of Hamiltonian<br />

mechanics that is exploited extensively for solv<strong>in</strong>g problems, as is described <strong>in</strong> chapter 15.<br />

It is advantageous to have the ability to exploit both the Lagrangian and Hamiltonian formulations simultaneously<br />

when handl<strong>in</strong>g systems that <strong>in</strong>volve a mixture of cyclic and non-cyclic coord<strong>in</strong>ates. The equations<br />

of motion for each <strong>in</strong>dependent generalized coord<strong>in</strong>ate can be derived <strong>in</strong>dependently of the rema<strong>in</strong><strong>in</strong>g generalized<br />

coord<strong>in</strong>ates. Thus it is possible to select either the Hamiltonian or the Lagrangian formulations for each<br />

generalized coord<strong>in</strong>ate, <strong>in</strong>dependent of what is used for the other generalized coord<strong>in</strong>ates. Routh[Rou1860]<br />

devised an elegant, and useful, hybrid technique that separates the cyclic and non-cyclic generalized coord<strong>in</strong>ates<br />

<strong>in</strong> order to simultaneously exploit the differ<strong>in</strong>g advantages of both the Hamiltonian and Lagrangian<br />

formulations of classical mechanics. The Routhian reduction approach partitions the P <br />

=1 ˙ k<strong>in</strong>etic energy<br />

term <strong>in</strong> the Hamiltonian <strong>in</strong>to a cyclic group, plus a non-cyclic group, i.e.<br />

( 1 ; 1 ; ) =<br />

X<br />

˙ − =<br />

=1<br />

X<br />

<br />

˙ +<br />

−<br />

X<br />

<br />

˙ − (8.58)<br />

Routh’s clever idea was to def<strong>in</strong>e a new function, called the Routhian, that <strong>in</strong>clude only one of the two<br />

partitions of the k<strong>in</strong>etic energy terms. This makes the Routhian a Hamiltonian for the coord<strong>in</strong>ates for which<br />

the k<strong>in</strong>etic energy terms are <strong>in</strong>cluded, while the Routhian acts like a negative Lagrangian for the coord<strong>in</strong>ates<br />

where the k<strong>in</strong>etic energy term is omitted. This book def<strong>in</strong>es two Routhians.<br />

( 1 ; ˙ 1 ˙ ; +1 ; )<br />

( 1 ; 1 ; ˙ +1 ˙ ; )<br />

≡<br />

≡<br />

X<br />

˙ − (8.59)<br />

<br />

X<br />

<br />

˙ − (8.60)<br />

The first, Routhian, called <strong>in</strong>cludes the k<strong>in</strong>etic energy terms only for the cyclic variables, and behaves<br />

like a Hamiltonian for the cyclic variables, and behaves like a Lagrangian for the non-cyclic variables. The<br />

second Routhian, called − <strong>in</strong>cludes the k<strong>in</strong>etic energy terms for only the non-cyclic variables, and<br />

behaves like a Hamiltonian for the non-cyclic variables, and behaves like a negative Lagrangian for the cyclic<br />

variables. These two Routhians complement each other <strong>in</strong> that they make the Routhian either a Hamiltonian<br />

for the cyclic variables, or the converse where the Routhian is a Hamiltonian for the non-cyclic variables.<br />

The Routhians use ( ˙ ) to denote those coord<strong>in</strong>ates for which the Routhian behaves like a Lagrangian, and<br />

( ) for those coord<strong>in</strong>ates where the Routhian behaves like a Hamiltonian. For uniformity, it is assumed<br />

that the degrees of freedom between 1 ≤ ≤ are non-cyclic, while those between +1 ≤ ≤ are ignorable<br />

cyclic coord<strong>in</strong>ates.<br />

The Routhian is a hybrid of Lagrangian and Hamiltonian mechanics. Some textbooks m<strong>in</strong>imize discussion<br />

of the Routhian on the grounds that this hybrid approach is not fundamental. However, the Routhian is<br />

used extensively <strong>in</strong> eng<strong>in</strong>eer<strong>in</strong>g <strong>in</strong> order to derive the equations of motion for rotat<strong>in</strong>g systems. In addition<br />

it is used when deal<strong>in</strong>g with rotat<strong>in</strong>g nuclei <strong>in</strong> nuclear physics, rotat<strong>in</strong>g molecules <strong>in</strong> molecular physics, and<br />

rotat<strong>in</strong>g galaxies <strong>in</strong> astrophysics. The Routhian reduction technique provides a powerful way to calculate<br />

the <strong>in</strong>tr<strong>in</strong>sic properties for a rotat<strong>in</strong>g system <strong>in</strong> the rotat<strong>in</strong>g frame of reference. The Routhian approach is<br />

<strong>in</strong>cluded <strong>in</strong> this textbook because it plays an important role <strong>in</strong> practical applications of rotat<strong>in</strong>g systems, plus<br />

it nicely illustrates the relative advantages of the Lagrangian and Hamiltonian formulations <strong>in</strong> mechanics.


8.6. ROUTHIAN REDUCTION 207<br />

8.6.1 R - Routhian is a Hamiltonian for the cyclic variables<br />

ThecyclicRouthian is def<strong>in</strong>ed assum<strong>in</strong>g that the variables between 1 ≤ ≤ are non-cyclic, where<br />

= −, while the variables between +1 ≤ ≤ are ignorable cyclic coord<strong>in</strong>ates. The cyclic Routhian<br />

expresses the cyclic coord<strong>in</strong>ates <strong>in</strong> terms of ( ) which are required for use by Hamilton’s equations,<br />

while the non-cyclic variables are expressed <strong>in</strong> terms of ( ˙) for use by the Lagrange equations. That is,<br />

the cyclic Routhian is def<strong>in</strong>ed to be<br />

X<br />

( 1 ; ˙ 1 ˙ ; +1 ; ) ≡ ˙ − (8.61)<br />

where the summation P ˙ is over only the cyclic variables +1 ≤ ≤ . Note that the Lagrangian<br />

can be split <strong>in</strong>to the cyclic and the non-cyclic parts<br />

X<br />

( 1 ; ˙ 1 ˙ ; +1 ; ) = ˙ − − (8.62)<br />

The first two terms on the right can be comb<strong>in</strong>ed to give the Hamiltonian for only the cyclic<br />

variables, = +1+2,thatis<br />

( 1 ; ˙ 1 ˙ ; +1 ; ) = − (8.63)<br />

The Routhian ( 1 ; ˙ 1 ˙ ; +1 ; ) also can be written <strong>in</strong> an alternate form<br />

X<br />

X<br />

X<br />

( 1 ; ˙ 1 ˙ ; +1 ; ) ≡ ˙ − = ˙ − − ˙ (8.64)<br />

<br />

= −<br />

<br />

X<br />

<br />

<br />

=1<br />

<br />

˙ (8.65)<br />

which is expressed as the complete Hamiltonian m<strong>in</strong>us the k<strong>in</strong>etic energy term for the noncyclic coord<strong>in</strong>ates.<br />

The Routhian behaves like a Hamiltonian for the cyclic coord<strong>in</strong>ates and behaves like a negative<br />

Lagrangian for all the = − noncyclic coord<strong>in</strong>ates =1 2 Thus the equations of motion<br />

for the non-cyclic variables are given us<strong>in</strong>g Lagrange’s equations of motion, while the Routhian behaves<br />

like a Hamiltonian for the ignorable cyclic variables = +1 <br />

Ignor<strong>in</strong>g both the Lagrange multiplier and generalized forces, then the partitioned equations of motion<br />

for the non-cyclic and cyclic generalized coord<strong>in</strong>ates are given <strong>in</strong> Table 81<br />

Table 81; Equations of motion for the Routhian <br />

Lagrange equations Hamilton equations<br />

Coord<strong>in</strong>ates Noncyclic: 1 ≤ ≤ Cyclic: ( +1)≤ ≤ <br />

Equations of motion<br />

<br />

<br />

= − <br />

<br />

<br />

<br />

= − ˙ <br />

<br />

˙ <br />

= − <br />

˙ <br />

<br />

<br />

= ˙ <br />

Thus there are cyclic (ignorable) coord<strong>in</strong>ates ( ) +1 ( ) <br />

which obey Hamilton’s equations of<br />

motion, while the the first = − non-cyclic (non-ignorable) coord<strong>in</strong>ates ( ˙) 1<br />

( ˙) <br />

for =1 2 <br />

obey Lagrange equations. The solution for the cyclic variables is trivial s<strong>in</strong>ce they are constants of motion<br />

and thus the Routhian has reduced the number of equations of motion that must be solved from to<br />

the = − non-cyclic variables This Routhian provides an especially useful way to reduce the number<br />

of equations of motion for rotat<strong>in</strong>g systems.<br />

Note that there are several def<strong>in</strong>itions used to def<strong>in</strong>e the Routhian, for example some books def<strong>in</strong>e this<br />

Routhian as be<strong>in</strong>g the negative of the def<strong>in</strong>ition used here so that it corresponds to a positive Lagrangian.<br />

However, this sign usually cancels when deriv<strong>in</strong>g the equations of motion, thus the sign convention is unimportant<br />

if a consistent sign convention is used.


208 CHAPTER 8. HAMILTONIAN MECHANICS<br />

8.6.2 R - Routhian is a Hamiltonian for the non-cyclic variables<br />

The non-cyclic Routhian complements . Aga<strong>in</strong> the generalized coord<strong>in</strong>ates between 1 ≤ ≤<br />

are assumed to be non-cyclic, while those between +1 ≤ ≤ are ignorable cyclic coord<strong>in</strong>ates. However,<br />

theexpression<strong>in</strong>termsof( ) and ( ˙) are <strong>in</strong>terchanged, that is, the cyclic variables are expressed <strong>in</strong><br />

terms of ( ˙) and the non-cyclic variables are expressed <strong>in</strong> terms of ( ) which is opposite of what was<br />

used for .<br />

( 1 ; 1 ; ˙ +1 ˙ ; ) =<br />

X<br />

<br />

˙ − − (8.66)<br />

= − (8.67)<br />

It can be written <strong>in</strong> a frequently used form<br />

( 1 ; 1 ; ˙ +1 ˙ ; )<br />

X<br />

X<br />

X<br />

≡ ˙ − = ˙ − − ˙ <br />

<br />

= −<br />

X<br />

<br />

=1<br />

<br />

˙ (8.68)<br />

This Routhian behaves like a Hamiltonian for the non-cyclic variables which are expressed <strong>in</strong> terms of <br />

and appropriate for a Hamiltonian. This Routhian writes the cyclic coord<strong>in</strong>ates <strong>in</strong> terms of , and ˙<br />

appropriate for a Lagrangian, which are treated assum<strong>in</strong>g the Routhian is a negative Lagrangian for<br />

these cyclic variables as summarized <strong>in</strong> table 82.<br />

Table 82; Equations of motion for the Routhian <br />

Hamilton equations Lagrange equations<br />

Coord<strong>in</strong>ates Noncyclic: 1 ≤ ≤ Cyclic: ( +1)≤ ≤ <br />

Equations of motion<br />

<br />

<br />

= − ˙ <br />

<br />

<br />

= − <br />

<br />

<br />

<br />

<br />

= ˙<br />

<br />

˙ <br />

= − <br />

˙ <br />

This non-cyclic Routhian is especially useful s<strong>in</strong>ce it equals the Hamiltonian for the non-cyclic<br />

variables, that is, the k<strong>in</strong>etic energy for motion of the cyclic variables has been removed. Note that s<strong>in</strong>ce the<br />

cyclic variables are constants of motion, then is a constant of motion if is a constant of motion.<br />

However, does not equal the total energy s<strong>in</strong>ce the coord<strong>in</strong>ate transformation is time dependent,<br />

that is, corresponds to the energy of the non-cyclic parts of the motion. For example, when used<br />

to describe rotational motion, corresponds to the energy <strong>in</strong> the non-<strong>in</strong>ertial rotat<strong>in</strong>g body-fixed<br />

frame of reference. This is especially useful <strong>in</strong> treat<strong>in</strong>g rotat<strong>in</strong>g systems such as rotat<strong>in</strong>g galaxies, rotat<strong>in</strong>g<br />

mach<strong>in</strong>ery, molecules, or rotat<strong>in</strong>g strongly-deformed nuclei as discussed <strong>in</strong> chapter 129<br />

The Lagrangian and Hamiltonian are the fundamental algebraic approaches to classical mechanics. The<br />

Routhian reduction method is a valuable hybrid technique that exploits a trick to reduce the number of<br />

variables that have to be solved for complicated problems encountered <strong>in</strong> science and eng<strong>in</strong>eer<strong>in</strong>g. The<br />

Routhian provides the most useful approach for solv<strong>in</strong>g the equations of motion for rotat<strong>in</strong>g<br />

molecules, deformed nuclei, or astrophysical objects <strong>in</strong> that it gives the Hamiltonian <strong>in</strong> the non-<strong>in</strong>ertial<br />

body-fixed rotat<strong>in</strong>g frame of reference ignor<strong>in</strong>g the rotational energy of the frame. By contrast, the cyclic<br />

Routhian is especially useful to exploit Lagrangian mechanics for solv<strong>in</strong>g problems <strong>in</strong> rigid-body<br />

rotation such as the Tippe Top described <strong>in</strong> example 1313.<br />

Note that the Lagrangian, Hamiltonian, plus both the and Routhian’s, all are scalars<br />

under rotation, that is, they are rotationally <strong>in</strong>variant. However, they may be expressed <strong>in</strong> terms of the<br />

coord<strong>in</strong>ates <strong>in</strong> either the stationary or a rotat<strong>in</strong>g frame. The major difference is that the Routhian <strong>in</strong>cludes<br />

only subsets of the k<strong>in</strong>etic energy term P ˙ . The relative merits of us<strong>in</strong>g Lagrangian, Hamiltonian, and<br />

both the and Routhian reduction methods, are illustrated by the follow<strong>in</strong>g examples.


8.6. ROUTHIAN REDUCTION 209<br />

8.6 Example: Spherical pendulum us<strong>in</strong>g Hamiltonian mechanics<br />

The spherical pendulum provides a simple test case for comparison<br />

of the use of Lagrangian mechanics, Hamiltonian mechanics,<br />

and both approaches to Routhian reduction. The Lagrangian mechanics<br />

solution of the spherical pendulum is described <strong>in</strong> example<br />

610. The solution us<strong>in</strong>g Hamiltonian mechanics is given <strong>in</strong> this<br />

example followed by solutions us<strong>in</strong>g both of the Routhian reduction<br />

approaches.<br />

Consider the equations of motion of a spherical pendulum of<br />

mass and length . The generalized coord<strong>in</strong>ates are s<strong>in</strong>ce<br />

the length is fixed at = The k<strong>in</strong>etic energy is<br />

g<br />

= 1 2 2 2 + 1 2 2 s<strong>in</strong> 2 2<br />

The potential energy = − cos giv<strong>in</strong>g that<br />

( ˙ ˙ ˙) = 1 2 2 2 + 1 2 2 s<strong>in</strong> 2 2 + cos <br />

m<br />

Spherical pendulum<br />

The generalized momenta are<br />

= <br />

˙ = 2 <br />

= <br />

˙ = 2 s<strong>in</strong> 2 <br />

S<strong>in</strong>ce the system is conservative, and the transformation from rectangular to spherical coord<strong>in</strong>ates does not<br />

depend explicitly on time, then the Hamiltonian is conserved and equals the total energy. The generalized<br />

momenta allow the Hamiltonian to be written as<br />

The equations of motion are<br />

( )=<br />

˙ = − <br />

=<br />

<br />

2 <br />

2 2 +<br />

2 <br />

2 2 s<strong>in</strong> 2 − cos <br />

<br />

2 cos <br />

2 2 s<strong>in</strong> 3 − s<strong>in</strong> <br />

()<br />

ṗ = − <br />

= 0<br />

˙ = = <br />

2<br />

˙ = <br />

=<br />

2 s<strong>in</strong> 2 <br />

Take the time derivative of equation () and use () to substitute for ˙ gives that<br />

¨ −<br />

2 cos <br />

2 4 s<strong>in</strong> 3 + s<strong>in</strong> =0<br />

()<br />

Note that equation (b) shows that is a cyclic coord<strong>in</strong>ate. Thus<br />

()<br />

()<br />

()<br />

= 2 s<strong>in</strong> 2 ˙ = constant<br />

that is the angular momentum about the vertical axis is conserved. Note that although is a constant of<br />

motion, ˙ <br />

=<br />

<br />

is a function of and thus <strong>in</strong> general it is not conserved. There are various solutions<br />

2 s<strong>in</strong> 2 <br />

depend<strong>in</strong>g on the <strong>in</strong>itial conditions. If =0then the pendulum is just the simple pendulum discussed<br />

previously that can oscillate, or rotate <strong>in</strong> the direction. The opposite extreme is where =0where the<br />

pendulum rotates <strong>in</strong> the direction with constant . In general the motion is a complicated coupl<strong>in</strong>g of the<br />

and motions.


210 CHAPTER 8. HAMILTONIAN MECHANICS<br />

8.7 Example: Spherical pendulum us<strong>in</strong>g ( ˙ ˙ )<br />

The Lagrangian for the spherical pendulum is<br />

( ˙ ˙ ˙) = 1 1<br />

2 2 ˙2 +<br />

2 2 s<strong>in</strong> 2 ˙ 2 + cos <br />

Note that the Lagrangian is <strong>in</strong>dependent of , therefore is an ignorable variable with<br />

˙ = <br />

= − =0<br />

Therefore is a constant of motion equal to<br />

The Routhian ( ˙ ˙ ) equals<br />

= <br />

˙ = 2 s<strong>in</strong> 2 ˙<br />

( ˙ ˙ ) = ˙ − <br />

∙ 1 1<br />

2¸<br />

= −<br />

2 2 ˙2 +<br />

2 2 s<strong>in</strong> 2 ˙ 2 + cos − 2 s<strong>in</strong> 2 ˙<br />

2 <br />

= − 1 1<br />

2 2 ˙2 +<br />

2 2 s<strong>in</strong> 2 + cos <br />

<br />

The Routhian ( ˙ ˙ ) behaves like a Hamiltonian for and like a Lagrangian 0 = − <br />

for . Use of Hamilton’s canonical equations for give<br />

˙ = <br />

=<br />

2 s<strong>in</strong> 2 <br />

− ˙ = <br />

=0<br />

<br />

These two equations show that is a constant of motion given by<br />

2 s<strong>in</strong> 2 ˙ = = constant<br />

()<br />

Note that the Hamiltonian only <strong>in</strong>cludes the k<strong>in</strong>etic energy for the motion which is a constant of motion,<br />

but this energy does not equal the total energy. This solution is what is predicted by Noether’s theorem due<br />

to the symmetry of the Lagrangian about the vertical axis.<br />

S<strong>in</strong>ce ( ˙ ˙ ) behaves like a Lagrangian for then the Lagrange equation for is<br />

Λ = <br />

˙<br />

− <br />

=0<br />

<br />

where the negative sign of the Lagrangian <strong>in</strong> ( ˙ ˙ ) cancels. This leads to<br />

that is<br />

2¨ 2 cos <br />

=<br />

2 s<strong>in</strong> 3 − s<strong>in</strong> <br />

<br />

¨ −<br />

2 cos <br />

2 4 s<strong>in</strong> 3 + s<strong>in</strong> =0<br />

()<br />

This result is identical to the one obta<strong>in</strong>ed us<strong>in</strong>g Lagrangian mechanics <strong>in</strong> example 610 and Hamiltonian<br />

mechanics given <strong>in</strong> example 86. The Routhian simplifiedtheproblemtoonedegreeoffreedom by<br />

absorb<strong>in</strong>g <strong>in</strong>to the Hamiltonian the ignorable cyclic coord<strong>in</strong>ate and its conserved conjugate momentum .<br />

Note that the central term <strong>in</strong> equation is the centrifugal term which is due to rotation about the vertical<br />

axis. This term is zero for plane pendulum motion when =0.


8.6. ROUTHIAN REDUCTION 211<br />

8.8 Example: Spherical pendulum us<strong>in</strong>g ( ˙)<br />

For a rotational system the Routhian ( ˙) also can be used to project out the Hamiltonian<br />

for the active variables <strong>in</strong> the rotat<strong>in</strong>g body-fixed frame of reference. Consider the spherical pendulum<br />

where the rotat<strong>in</strong>g frame is rotat<strong>in</strong>g with angular velocity ˙. The Lagrangian for the spherical pendulum is<br />

( ˙ ˙ ˙) = 1 1<br />

2 2 ˙2 +<br />

2 2 s<strong>in</strong> 2 ˙ 2 + cos <br />

Note that the Lagrangian is <strong>in</strong>dependent of , therefore is an ignorable variable with<br />

˙ = <br />

= − =0<br />

Therefore is a constant of motion equal to<br />

= <br />

˙ = 2 s<strong>in</strong> 2 ˙<br />

The total Hamiltonian is given by<br />

( )= X ˙ − =<br />

2 <br />

2 2 +<br />

2 <br />

2 2 s<strong>in</strong> 2 − cos <br />

<br />

<br />

The Routhian for the rotat<strong>in</strong>g frame of reference is given by equation 868, that is<br />

( ˙)<br />

X<br />

= ˙ − ˙ − = − ˙<br />

=1<br />

=<br />

2 <br />

2 2 +<br />

2 <br />

2 <br />

2 2 s<strong>in</strong> 2 − cos − ˙<br />

=<br />

2 2 − 1 2 2 s<strong>in</strong> 2 ˙ 2 − cos <br />

This behaves like a negative Lagrangian for and a Hamiltonian for . The conjugate momenta are<br />

= <br />

˙ = − <br />

˙<br />

= 2 s<strong>in</strong> 2 ˙<br />

˙ = <br />

= − <br />

=0<br />

<br />

that is, is a constant of motion.<br />

Hamilton’s equations of motion give<br />

()<br />

˙ = <br />

<br />

= <br />

2<br />

− ˙ = <br />

<br />

= − 2 cos <br />

2 s<strong>in</strong> 3 + s<strong>in</strong> <br />

()<br />

Equation gives that<br />

<br />

˙ = ¨ = ˙ <br />

2<br />

Insert<strong>in</strong>g this <strong>in</strong>to equation gives<br />

¨ −<br />

2 cos <br />

2 4 s<strong>in</strong> 3 + s<strong>in</strong> =0<br />

<br />

which is identical to the equation of motion derived us<strong>in</strong>g .TheHamiltonian<strong>in</strong>therotat<strong>in</strong>gframe<br />

is a constant of motion given by but it does not <strong>in</strong>clude the total energy.<br />

Note that these examples show that both forms of the Routhian, as well as the complete Lagrangian<br />

formalism, shown <strong>in</strong> example 610, and complete Hamiltonian formalism, shown <strong>in</strong> example 86 all give the<br />

same equations of motion. This illustrates that the Lagrangian, Hamiltonian, and Routhian mechanics all<br />

give the same equations of motion and this applies both <strong>in</strong> the static <strong>in</strong>ertial frame as well as a rotat<strong>in</strong>g frame<br />

s<strong>in</strong>ce the Lagrangian, Hamiltonian and Routhian all are scalars under rotation, that is, they are rotationally<br />

<strong>in</strong>variant.<br />

()


212 CHAPTER 8. HAMILTONIAN MECHANICS<br />

8.9 Example: S<strong>in</strong>gle particle mov<strong>in</strong>g <strong>in</strong> a vertical plane under the <strong>in</strong>fluence of<br />

an <strong>in</strong>verse-square central force<br />

TheLagrangianforas<strong>in</strong>gleparticleofmass mov<strong>in</strong>g <strong>in</strong> a vertical plane and subject to a central <strong>in</strong>verse<br />

square central force, is specified by two generalized coord<strong>in</strong>ates, and <br />

= 2 ( ˙2 + 2 ˙2 )+<br />

<br />

<br />

The ignorable coord<strong>in</strong>ate is s<strong>in</strong>ce it is cyclic. Let the constant conjugate momentum be denoted by =<br />

<br />

˙ = 2 ˙. Then the correspond<strong>in</strong>g cyclic Routhian is<br />

( ˙ )= ˙ − =<br />

2 <br />

2 2 − 1 2 ˙2 − <br />

This Routhian is the equivalent one-dimensional potential () m<strong>in</strong>us the k<strong>in</strong>etic energy of radial motion.<br />

Apply<strong>in</strong>g Hamilton’s equation to the cyclic coord<strong>in</strong>ate gives<br />

˙ =0<br />

<br />

2 = ˙<br />

imply<strong>in</strong>g a solution<br />

= 2 ˙ = <br />

where the angular momentum is a constant.<br />

The Lagrange-Euler equation can be applied to the non-cyclic coord<strong>in</strong>ate <br />

Λ = <br />

− <br />

=0<br />

˙ <br />

where the negative sign of cancels. This leads to the radial solution<br />

¨ −<br />

2 <br />

3 + 2 =0<br />

where = which is a constant of motion <strong>in</strong> the centrifugal term. Thus the problem has been reduced to a<br />

one-dimensional problem <strong>in</strong> radius that is <strong>in</strong> a rotat<strong>in</strong>g frame of reference.<br />

8.7 Variable-mass systems<br />

Lagrangian and Hamiltonian mechanics assume that the total mass and energy of the system are conserved.<br />

Variable-mass systems <strong>in</strong>volve transferr<strong>in</strong>g mass and energy between donor and receptor bodies. However,<br />

such systems still can be conservative if the Lagrangian or Hamiltonian <strong>in</strong>clude all the active degrees of<br />

freedom for the comb<strong>in</strong>ed donor-receptor system. The follow<strong>in</strong>g examples of variable mass systems illustrate<br />

subtle complications that occur handl<strong>in</strong>g such problems us<strong>in</strong>g algebraic mechanics.<br />

8.7.1 Rocket propulsion:<br />

Newtonian mechanics was used to solve the rocket problem <strong>in</strong> chapter 2126. The equation of motion<br />

(2113) relat<strong>in</strong>g the rocket thrust to the rate of change of the momentum separated <strong>in</strong>to two terms,<br />

= ˙ = ¨ + ˙ ˙ (8.69)<br />

The first term is the usual mass times acceleration, while the second term arises from the rate of change of<br />

mass times the velocity. The equation of motion for rocket motion is easily derived us<strong>in</strong>g either Lagrangian<br />

or Hamiltonian mechanics by relat<strong>in</strong>g the rocket thrust to the generalized force


8.7. VARIABLE-MASS SYSTEMS 213<br />

8.7.2 Mov<strong>in</strong>g cha<strong>in</strong>s:<br />

The motion of a flexible, frictionless, heavy cha<strong>in</strong> that is fall<strong>in</strong>g <strong>in</strong> a gravitational field, often can be split <strong>in</strong>to<br />

two coupled variable-mass partitions that have different cha<strong>in</strong>-l<strong>in</strong>k velocities. These partitions are coupled<br />

at the mov<strong>in</strong>g <strong>in</strong>tersection between the cha<strong>in</strong> partitions. That is, these partitions share time-dependent<br />

fractions of the total cha<strong>in</strong> mass. Mov<strong>in</strong>g cha<strong>in</strong>s were discussed first by Caley <strong>in</strong> 1857[Cay1857] and s<strong>in</strong>ce<br />

then the mov<strong>in</strong>g cha<strong>in</strong> problem has had a controversial history due to the frequent erroneous assumption<br />

that, <strong>in</strong> the gravitational field, the cha<strong>in</strong> partitions fall with acceleration rather than apply<strong>in</strong>g the correct<br />

energy conservation assumption for this conservative system. The follow<strong>in</strong>g two examples of conservative<br />

fall<strong>in</strong>g-cha<strong>in</strong> systems illustrate solutions obta<strong>in</strong>ed us<strong>in</strong>g variational pr<strong>in</strong>ciples applied to a s<strong>in</strong>gle cha<strong>in</strong> that<br />

is partitioned <strong>in</strong>to two variable length sections. 1<br />

Consider the follow<strong>in</strong>g two possible scenarios for motion of a flexible, heavy, frictionless, cha<strong>in</strong> located <strong>in</strong><br />

a uniform gravitational field . Thefirst scenario is the “folded cha<strong>in</strong>” system which assumes that one end<br />

of the cha<strong>in</strong> is held fixed, while the adjacent free end is released at the same altitude as the top of the fixed<br />

arm, and this free end is allowed to fall <strong>in</strong> the constant gravitational field . The second “fall<strong>in</strong>g cha<strong>in</strong>”,<br />

scenario assumes that one end of the cha<strong>in</strong> is hang<strong>in</strong>g down through a hole <strong>in</strong> a frictionless, smooth, rigid,<br />

horizontal table, with the stationary partition of the cha<strong>in</strong> sitt<strong>in</strong>g on the table surround<strong>in</strong>g the hole. The<br />

fall<strong>in</strong>g section of this cha<strong>in</strong> is be<strong>in</strong>g pulled out of the stationary pile by the hang<strong>in</strong>g partition. Both of these<br />

systems are conservative s<strong>in</strong>ce it is assumed that the total mass of the cha<strong>in</strong> is fixed, and no dissipative forces<br />

are act<strong>in</strong>g. The cha<strong>in</strong>s are assumed to be <strong>in</strong>extensible, flexible, and frictionless, and subject to a uniform<br />

gravitational field <strong>in</strong> the vertical direction. In both examples, the cha<strong>in</strong>, with mass and length is<br />

partitioned <strong>in</strong>to a stationary segment, plus a mov<strong>in</strong>g segment, where the mass per unit length of the cha<strong>in</strong><br />

is = <br />

. These partitions are strongly coupled at their <strong>in</strong>tersection which propagates downward with time<br />

for the “folded cha<strong>in</strong>” and propagates upward, relative to the lower end of the fall<strong>in</strong>g cha<strong>in</strong>, for the “fall<strong>in</strong>g<br />

cha<strong>in</strong>”. For the “folded cha<strong>in</strong>”, the cha<strong>in</strong> l<strong>in</strong>ks are transferred from the mov<strong>in</strong>g segment to the stationary<br />

segment as the mov<strong>in</strong>g section falls. By contrast, for the “fall<strong>in</strong>g system”, the cha<strong>in</strong> l<strong>in</strong>ks are transferred<br />

from the stationary upper section to the mov<strong>in</strong>g lower segment of the cha<strong>in</strong>.<br />

8.10 Example: Folded cha<strong>in</strong><br />

The folded cha<strong>in</strong> of length and mass-per-unit-length = <br />

hangs<br />

vertically downwards <strong>in</strong> a gravitational field with both ends held <strong>in</strong>itially<br />

at the same height. The fixed end is attached to a fixed support while the<br />

free end of the cha<strong>in</strong> is dropped at time =0withthefreeendatthesame<br />

height and adjacent to the fixed end. Let be the distance the fall<strong>in</strong>g free<br />

end is below the fixed end. Us<strong>in</strong>g an idealized one-dimensional assumption,<br />

the Lagrangian L is given by<br />

L( ˙) = 4 ( − ) ˙2 + 1<br />

4 (2 +2 − 2 ) (8.70)<br />

where the bracket <strong>in</strong> the second term is the height of the center of mass of<br />

the folded cha<strong>in</strong> with respect to the fixed upper end of the cha<strong>in</strong>.<br />

The Hamiltonian is given by<br />

L + y<br />

2<br />

y<br />

L - y<br />

2<br />

<br />

( )= ˙ − L( ˙) =<br />

<br />

( − ) − +2 − 2 )<br />

(2 (8.71)<br />

4<br />

where is the l<strong>in</strong>ear momentum of the right-hand arm of the folded cha<strong>in</strong>.<br />

As shown <strong>in</strong> the discussion of the Generalized Energy Theorem, (chapters 78 and 79), when all the<br />

active forces are <strong>in</strong>cluded <strong>in</strong> the Lagrangian and the Hamiltonian, then the total mechanical energy is<br />

given by = Moreover, both the Lagrangian and the Hamiltonian are time <strong>in</strong>dependent, s<strong>in</strong>ce<br />

<br />

= = −L =0 (8.72)<br />

<br />

Therefore the “folded cha<strong>in</strong>” Hamiltonian equals the total energy, which is a constant of motion. Energy<br />

conservation for this system can be used to give<br />

1 Discussions with Professor Frank Wolfs stimulated <strong>in</strong>clusion of these two examples of mov<strong>in</strong>g cha<strong>in</strong>s.


214 CHAPTER 8. HAMILTONIAN MECHANICS<br />

<br />

4 ( − ) ˙2 − 1 4 (2 +2 − 2 )=− 1 4 2 (8.73)<br />

Solve for ˙ 2 gives<br />

˙ 2 = (2 − 2 )<br />

(8.74)<br />

− <br />

The acceleration of the fall<strong>in</strong>g arm, ¨ is given by tak<strong>in</strong>g the time derivative of equation 874<br />

¨ = + ¡ 2 − 2¢<br />

(8.75)<br />

2( − )<br />

The rate of change <strong>in</strong> l<strong>in</strong>ear momentum for the mov<strong>in</strong>g right side of the cha<strong>in</strong>, ˙ , is given by<br />

˙ = ¨ + ˙ ˙ = + (2 − 2 )<br />

2( − )<br />

(8.76)<br />

For this energy-conserv<strong>in</strong>g cha<strong>in</strong>, the tension <strong>in</strong> the cha<strong>in</strong> 0 at the fixed end of the cha<strong>in</strong> is given by<br />

0 = <br />

2 ( + )+1 4 ˙2 (8.77)<br />

Equations 874 and 876, imply that the tension diverges to <strong>in</strong>f<strong>in</strong>ity when → . Calk<strong>in</strong> and March<br />

measured the dependence of the cha<strong>in</strong> tension at the support for the folded cha<strong>in</strong> and observed the predicted<br />

dependence. The maximum tension was ' 25 which is consistent with that predicted us<strong>in</strong>g equation 877<br />

after tak<strong>in</strong>g <strong>in</strong>to account the f<strong>in</strong>ite size and mass of <strong>in</strong>dividual l<strong>in</strong>ks <strong>in</strong> the cha<strong>in</strong>. This result is very different<br />

from that obta<strong>in</strong>ed us<strong>in</strong>g the erroneous assumption that the right arm falls with the free-fall acceleration ,<br />

which implies a maximum tension 0 =2. Thus the free-fall assumption disagrees with the experimental<br />

results, <strong>in</strong> addition to violat<strong>in</strong>g energy conservation and the tenets of Lagrangian and Hamiltonian mechanics.<br />

That is, the experimental result demonstrates unambiguously that the energy conservation predictions apply<br />

<strong>in</strong> contradiction with the erroneous free-fall assumption.<br />

The unusual feature of variable mass problems, such as the folded cha<strong>in</strong> problem, is that the rate of change<br />

of momentum <strong>in</strong> equation 876 <strong>in</strong>cludes two contributions to the force and rate of change of momentum,<br />

that is, it <strong>in</strong>cludes both the acceleration term ¨ plus the variable mass term ˙ ˙ that accounts for the<br />

transfer of matter at the <strong>in</strong>tersection of the mov<strong>in</strong>g and stationary partitions of the cha<strong>in</strong>. At the transition<br />

po<strong>in</strong>t of the cha<strong>in</strong>, mov<strong>in</strong>g l<strong>in</strong>ks are transferred from the mov<strong>in</strong>g section and are added to the stationary<br />

subsection. S<strong>in</strong>ce this mov<strong>in</strong>g section is fall<strong>in</strong>g downwards, and the stationary section is stationary, then the<br />

transferred momentum is <strong>in</strong> a downward direction correspond<strong>in</strong>g to an <strong>in</strong>creased effective downward force.<br />

Thus the measured acceleration of the mov<strong>in</strong>g arm actually is faster than . A related phenomenon is the<br />

loud crack<strong>in</strong>g sound heard when crack<strong>in</strong>g a whip.<br />

8.11 Example: Fall<strong>in</strong>g cha<strong>in</strong><br />

The “fall<strong>in</strong>g cha<strong>in</strong>”, scenario assumes that one end of the cha<strong>in</strong> is hang<strong>in</strong>g<br />

down through a hole <strong>in</strong> a frictionless, smooth, rigid, horizontal table,<br />

with the stationary partition of the cha<strong>in</strong> ly<strong>in</strong>g on the frictionless table surround<strong>in</strong>g<br />

the hole. The fall<strong>in</strong>g section of this cha<strong>in</strong> is be<strong>in</strong>g pulled out of<br />

the stationary pile by the hang<strong>in</strong>g partition. The analysis for the problem of<br />

the fall<strong>in</strong>g cha<strong>in</strong> behaves differently from the folded cha<strong>in</strong>. For the “fall<strong>in</strong>gcha<strong>in</strong>”<br />

let be the fall<strong>in</strong>g distance of the lower end of the cha<strong>in</strong> measured<br />

with respect to the table top. The Lagrangian and Hamiltonian are given by<br />

y<br />

L( ˙) = 2 ˙2 + 2<br />

2<br />

(8.78)<br />

= L = ˙<br />

˙<br />

(8.79)<br />

= 2 <br />

2 − 2 = (8.80)<br />

2


8.8. SUMMARY 215<br />

The Lagrangian and Hamiltonian are not explicitly time dependent, and the Hamiltonian equals the <strong>in</strong>itial<br />

total energy, 0 . Thus energy conservation can be used to give that<br />

Lagrange’s equation of motion gives<br />

= 1 2 ( ˙2 − ) = 0 (8.81)<br />

˙ = ¨ + ˙ ˙ = + 1 2 ˙2 = − 0 (8.82)<br />

The important difference between the folded cha<strong>in</strong> and fall<strong>in</strong>g cha<strong>in</strong> is that the mov<strong>in</strong>g component of the<br />

fall<strong>in</strong>g cha<strong>in</strong> is ga<strong>in</strong><strong>in</strong>g mass with time rather than los<strong>in</strong>g mass. Also the tension <strong>in</strong> the cha<strong>in</strong> 0 reduces the<br />

acceleration of the fall<strong>in</strong>g cha<strong>in</strong> mak<strong>in</strong>g it less than the free-fall value . This is <strong>in</strong> contrast to that for the<br />

folded cha<strong>in</strong> system where the acceleration exceeds .<br />

The above discussion shows that Lagrangian and Hamiltonian can be applied to variable-mass systems if<br />

both the donor and receptor degrees of freedom are <strong>in</strong>cluded to ensure that the total mass is conserved.<br />

8.8 Summary<br />

Hamilton’s equations of motion<br />

Insert<strong>in</strong>g the generalized momentum <strong>in</strong>to Jacobi’s generalized energy relation was used to def<strong>in</strong>e the<br />

Hamiltonian function to be<br />

(q p)=p · ˙q−(q ˙q)<br />

(83)<br />

The Legendre transform of the Lagrange-Euler equations, led to Hamilton’s equations of motion.<br />

˙ = <br />

(825)<br />

<br />

˙ = − <br />

<br />

+<br />

" <br />

X<br />

=1<br />

<br />

+ <br />

<br />

<br />

The generalized energy equation 738 gives the time dependence<br />

(q p)<br />

= X Ã" <br />

# !<br />

X <br />

+ <br />

˙ −<br />

<br />

<br />

<br />

<br />

=1<br />

#<br />

(q ˙q)<br />

<br />

(826)<br />

(827)<br />

where<br />

<br />

= −<br />

(824)<br />

<br />

The are treated as <strong>in</strong>dependent canonical variables Lagrange was the first to derive the canonical<br />

equations but he did not recognize them as a basic set of equations of motion. Hamilton derived the canonical<br />

equations of motion from his fundamental variational pr<strong>in</strong>ciple and made them the basis for a far-reach<strong>in</strong>g<br />

theory of dynamics. Hamilton’s equations give 2 first-order differential equations for for each of the<br />

degrees of freedom. Lagrange’s equations give second-order differential equations for the variables ˙ <br />

Routhian reduction technique<br />

The Routhian reduction technique is a hybrid of Lagrangian and Hamiltonian mechanics that exploits<br />

the advantages of both approaches for solv<strong>in</strong>g problems <strong>in</strong>volv<strong>in</strong>g cyclic variables. It is especially useful for<br />

solv<strong>in</strong>g motion <strong>in</strong> rotat<strong>in</strong>g systems <strong>in</strong> science and eng<strong>in</strong>eer<strong>in</strong>g. Two Routhians are used frequently for solv<strong>in</strong>g<br />

the equations of motion of rotat<strong>in</strong>g systems. Assum<strong>in</strong>g that the variables between 1 ≤ ≤ are non-cyclic,<br />

while the variables between +1≤ ≤ are ignorable cyclic coord<strong>in</strong>ates, then the two Routhians are:<br />

( 1 ; ˙ 1 ˙ ; +1 ; ) =<br />

( 1 ; 1 ; ˙ +1 ˙ ; ) =<br />

X<br />

<br />

X<br />

<br />

˙ − = −<br />

˙ − = −<br />

X<br />

<br />

X<br />

<br />

˙ <br />

˙ <br />

(865)<br />

(868)


216 CHAPTER 8. HAMILTONIAN MECHANICS<br />

The Routhian is a negative Lagrangian for the non-cyclic variables between 1 ≤ ≤ , where<br />

= − and is a Hamiltonian for the cyclic variables between +1 ≤ ≤ . S<strong>in</strong>ce the cyclic<br />

variables are constants of the Hamiltonian, their solution is trivial, and the number of variables <strong>in</strong>cluded <strong>in</strong><br />

the Lagrangian is reduced from to = − . The Routhian is useful for solv<strong>in</strong>g some problems <strong>in</strong><br />

classical mechanics. The Routhian is a Hamiltonian for the non-cyclic variables between 1 ≤ ≤ ,<br />

and is a negative Lagrangian for the cyclic variables between +1≤ ≤ . S<strong>in</strong>cethecyclicvariables<br />

are constants of motion, the Routhian also is a constant of motion but it does not equal the total<br />

energy s<strong>in</strong>ce the coord<strong>in</strong>ate transformation is time dependent. The Routhian is especially valuable<br />

for solv<strong>in</strong>g rotat<strong>in</strong>g many-body systems such as galaxies, molecules, or nuclei, s<strong>in</strong>ce the Routhian <br />

is the Hamiltonian <strong>in</strong> the rotat<strong>in</strong>g body-fixed coord<strong>in</strong>ate frame.<br />

Variable mass systems:<br />

Two examples of heavy flexible cha<strong>in</strong>s fall<strong>in</strong>g <strong>in</strong> a uniform gravitational field were used to illustrate<br />

how variable mass systems can be handled us<strong>in</strong>g Lagrangian and Hamiltonian mechanics. The fall<strong>in</strong>g-mass<br />

system is conservative assum<strong>in</strong>g that both the donor plus the receptor body systems are <strong>in</strong>cluded.<br />

Comparison of Lagrangian and Hamiltonian mechanics<br />

Lagrangian and the Hamiltonian dynamics are two powerful and related variational algebraic formulations<br />

of mechanics that are based on Hamilton’s action pr<strong>in</strong>ciple. They can be applied to any conservative degrees<br />

of freedom as discussed <strong>in</strong> chapters 6 8 and 15. Lagrangian and Hamiltonian mechanics both concentrate<br />

solely on active forces and can ignore <strong>in</strong>ternal forces. They can handle many-body systems and allow<br />

convenient generalized coord<strong>in</strong>ates of choice. This ability is impractical or impossible us<strong>in</strong>g Newtonian<br />

mechanics. Thus it is natural to compare the relative advantages of these two algebraic formalisms <strong>in</strong> order<br />

to decide which should be used for a specific problem.<br />

For a system with generalized coord<strong>in</strong>ates, plus constra<strong>in</strong>t forces that are not required to be known,<br />

then the Lagrangian approach, us<strong>in</strong>g a m<strong>in</strong>imal set of generalized coord<strong>in</strong>ates, reduces to only = − <br />

second-order differential equations and unknowns compared to the Newtonian approach where there are<br />

+ unknowns. Alternatively, use of Lagrange multipliers allows determ<strong>in</strong>ation of the constra<strong>in</strong>t forces<br />

result<strong>in</strong>g <strong>in</strong> + second order equations and unknowns. The Lagrangian potential function is limited<br />

to conservative forces, Lagrange multipliers can be used to handle holonomic forces of constra<strong>in</strong>t, while<br />

generalized forces can be used to handle non-conservative and non-holonomic forces. The advantage of the<br />

Lagrange equations of motion is that they can deal with any type of force, conservative or non-conservative,<br />

and they directly determ<strong>in</strong>e , ˙ rather than whichthenrequiresrelat<strong>in</strong>g to ˙.<br />

For a system with generalized coord<strong>in</strong>ates, the Hamiltonian approach determ<strong>in</strong>es 2 first-order differential<br />

equations which are easier to solve than second-order equations. However, the 2 solutions must be<br />

comb<strong>in</strong>ed to determ<strong>in</strong>e the equations of motion. The Hamiltonian approach is superior to the Lagrange approach<br />

<strong>in</strong> its ability to obta<strong>in</strong> an analytical solution of the <strong>in</strong>tegrals of the motion. Hamiltonian dynamics also<br />

has a means of determ<strong>in</strong><strong>in</strong>g the unknown variables for which the solution assumes a soluble form. Important<br />

applications of Hamiltonian mechanics are to quantum mechanics and statistical mechanics, where quantum<br />

analogs of and can be used to relate to the fundamental variables of Hamiltonian mechanics. This<br />

does not apply for the variables and ˙ of Lagrangian mechanics. The Hamiltonian approach is especially<br />

powerful when the system has cyclic variables, then the conjugate momenta are constants. Thus the<br />

conjugate variables ( ) can be factored out of the Hamiltonian, which reduces the number of conjugate<br />

variables required to − . This is not possible us<strong>in</strong>g the Lagrangian approach s<strong>in</strong>ce, even though the <br />

coord<strong>in</strong>ates canbefactoredout,thevelocities ˙ still must be <strong>in</strong>cluded, thus the conjugate variables<br />

must be <strong>in</strong>cluded. The Lagrange approach is advantageous for obta<strong>in</strong><strong>in</strong>g a numerical solution of systems <strong>in</strong><br />

classical mechanics. However, Hamiltonian mechanics expresses the variables <strong>in</strong> terms of the fundamental<br />

canonical variables (q p) which provides a more fundamental <strong>in</strong>sight <strong>in</strong>to the underly<strong>in</strong>g physics. 2<br />

2 Recommended read<strong>in</strong>g: "<strong>Classical</strong> <strong>Mechanics</strong>" H. Goldste<strong>in</strong>, Addison-Wesley, Read<strong>in</strong>g (1950). The present chapter<br />

closely follows the notation used by Goldste<strong>in</strong> to facilitate cross-referenc<strong>in</strong>g and read<strong>in</strong>g the many other textbooks that have<br />

adopted this notation.


8.8. SUMMARY 217<br />

Problems<br />

1. Ablockofmass rests on an <strong>in</strong>cl<strong>in</strong>ed plane mak<strong>in</strong>g an angle with the horizontal. The <strong>in</strong>cl<strong>in</strong>ed plane (a<br />

triangular block of mass ) is free to slide horizontally without friction. The block of mass is also free to<br />

slide on the larger block of mass without friction.<br />

(a) Construct the Lagrangian function.<br />

(b) Derive the equations of motion for this system.<br />

(c) Calculate the canonical momenta.<br />

(d) Construct the Hamiltonian function.<br />

(e) F<strong>in</strong>d which of the two momenta found <strong>in</strong> part (c) is a constant of motion and discuss why it is so. If the<br />

two blocks start from rest, what is the value of this constant of motion?<br />

2. Discuss among yourselves the follow<strong>in</strong>g four conditions that can exist for the Hamiltonian and give several<br />

examples of systems exhibit<strong>in</strong>g each of the four conditions.<br />

(a) The Hamiltonian is conserved and equals the total mechanical energy<br />

(b) The Hamiltonian is conserved but does not equal the total mechanical energy<br />

(c) The Hamiltonian is not conserved but does equal the total mechanical energy<br />

(d) The Hamiltonian is not conserved and does not equal the mechanical total energy.<br />

3. Ablockofmass rests on an <strong>in</strong>cl<strong>in</strong>ed plane mak<strong>in</strong>g an angle with the horizontal. The <strong>in</strong>cl<strong>in</strong>ed plane (a<br />

triangular block of mass ) is free to slide horizontally without friction. The block of mass is also free to<br />

slide on the larger block of mass without friction.<br />

(a) Construct the Lagrangian function.<br />

(b) Derive the equations of motion for this system.<br />

(c) Calculate the canonical momenta.<br />

(d) Construct the Hamiltonian function.<br />

(e) F<strong>in</strong>d which of the two momenta found <strong>in</strong> part (c) is a constant of motion and discuss why it is so. If the<br />

two blocks start from rest, what is the value of this constant of motion?<br />

4. Discuss among yourselves the follow<strong>in</strong>g four conditions that can exist for the Hamiltonian and give several<br />

examples of systems exhibit<strong>in</strong>g each of the four conditions.<br />

a) The Hamiltonian is conserved and equals the total mechanical energy<br />

b) The Hamiltonian is conserved but does not equal the total mechanical energy<br />

c) The Hamiltonian is not conserved but does equal the total mechanical energy<br />

d) The Hamiltonian is not conserved and does not equal the mechanical total energy<br />

5. Compare the Lagrangian formalism and the Hamiltonian formalism by creat<strong>in</strong>g a two-column chart. Label one<br />

side “Lagrangian” and the other side “Hamiltonian” and discuss the similarities and differences. Here are some<br />

ideastogetyoustarted:<br />

• What are the basic variables <strong>in</strong> each formalism?<br />

• What are the form and number of the equations of motion derived <strong>in</strong> each case?<br />

• How does the Lagrangian “state space” compare to the Hamiltonian “phase space”?<br />

6. It can be shown that if ( ˙ ) is the Lagrangian of a particle mov<strong>in</strong>g <strong>in</strong> one dimension, then = 0 where<br />

0 ( ˙ ) =( ˙ ) + <br />

<br />

and ( ) is an arbitrary function. This problem explores the consequences of<br />

this on the Hamiltonian formalism.


218 CHAPTER 8. HAMILTONIAN MECHANICS<br />

(a) Relate the new canonical momentum 0 ,for 0 , to the old canonical momentum , for.<br />

(b) Express the new Hamiltonian 0 ( 0 0 ) for 0 <strong>in</strong> terms of the old Hamiltonian ( ) and .<br />

(c) Explicitly show that the new Hamilton’s equations for 0 are equivalent to the old Hamilton’s equations<br />

for .<br />

7. A massless hoop of radius is rotat<strong>in</strong>g about an axis perpendicular to its central axis at constant angular<br />

velocity . Amass can freely slide around the hoop.<br />

(a) Determ<strong>in</strong>e the Lagrangian of the system.<br />

(b) Determ<strong>in</strong>e the Hamiltonian of the system. Does it equal the total mechanical energy?<br />

(c) Determ<strong>in</strong>e the Lagrangian of the system with respect to a coord<strong>in</strong>ate frame <strong>in</strong> which = + eff .What<br />

is eff ? What force generates the additional term <strong>in</strong> eff ?<br />

8. Consider a pendulum of length attached to the end of rod of length . The rod is rotat<strong>in</strong>g at constant<br />

angular velocity <strong>in</strong> the plane. Assume the pendulum is always taut.<br />

(a) Determ<strong>in</strong>e equations of motion.<br />

(b) For what value of 2 is this system the same as a plane pendulum <strong>in</strong> a constant gravitational field?<br />

(c) Show 6= . What is the reason?<br />

9. Aparticleofmass <strong>in</strong> a gravitational field slides on the <strong>in</strong>side of a smooth parabola of revolution whose axis<br />

is vertical. Us<strong>in</strong>g the distance from the axis and the azimuthal angle as generalized coord<strong>in</strong>ates, f<strong>in</strong>d the<br />

follow<strong>in</strong>g.<br />

(a) The Lagrangian of the system.<br />

(b) The generalized momenta and the correspond<strong>in</strong>g Hamiltonian<br />

(c) The equation of motion for the coord<strong>in</strong>ate as a function of time.<br />

(d) If <br />

<br />

=0 show that the particle can execute small oscillations about the lowest po<strong>in</strong>t of the paraboloid<br />

and f<strong>in</strong>d the frequency of these oscillations.<br />

10. Consider a particle of mass which is constra<strong>in</strong>ed to move on the surface of a sphere of radius . Thereare<br />

no external forces of any k<strong>in</strong>d act<strong>in</strong>g on the particle.<br />

(a) What is the number of generalized coord<strong>in</strong>ates necessary to describe the problem?<br />

(b) Choose a set of generalized coord<strong>in</strong>ates and write the Lagrangian of the system.<br />

(c) What is the Hamiltonian of the system? Is it conserved?<br />

(d) Prove that the motion of the particle is along a great circle of the sphere.<br />

11. Ablockofmass is attached to a wedge of mass byaspr<strong>in</strong>gwithspr<strong>in</strong>gconstant. The <strong>in</strong>cl<strong>in</strong>ed<br />

frictionless surface of the wedge makes an angle to the horizontal. The wedge is free to slide on a horizontal<br />

frictionless surface as shown <strong>in</strong> the figure.<br />

(a) Giventhattherelaxedlengthofthespr<strong>in</strong>gis, f<strong>in</strong>d the values 0 when both book and wedge are<br />

stationary.<br />

(b) F<strong>in</strong>d the Lagrangian for the system as a function of the coord<strong>in</strong>ate of the wedge and the length of<br />

spr<strong>in</strong>g . Write down the equations of motion.<br />

(c) What is the natural frequency of vibration?


8.8. SUMMARY 219<br />

12. A fly-ball governor comprises two masses connected by 4 h<strong>in</strong>ged arms of length toaverticalshaftandtoa<br />

mass which can slide up or down the shaft without friction <strong>in</strong> a uniform vertical gravitational field as shown<br />

<strong>in</strong> the figure. The assembly is constra<strong>in</strong>ed to rotate around the axis of the vertical shaft with same angular<br />

velocity as that of the vertical shaft. Neglect the mass of the arms, air friction, and assume that the mass <br />

has a negligible moment of <strong>in</strong>ertia. Assume that the whole system is constra<strong>in</strong>ed to rotate with a constant<br />

angular velocity 0 .<br />

(a) Choose suitable coord<strong>in</strong>ates and use the Lagrangian to derive equations of motion of the system around<br />

the equilibrium position.<br />

(b) Determ<strong>in</strong>e the height of the mass above its lowest position as a function of 0 .<br />

(c) F<strong>in</strong>d the frequency of small oscillations about this steady motion.<br />

(d) Derive a Routhian that provides the Hamiltonian <strong>in</strong> the rotat<strong>in</strong>g system.<br />

(e) Is the total energy of the fly-ball governor <strong>in</strong> the rotat<strong>in</strong>g frame of reference constant <strong>in</strong> time?<br />

(f) Suppose that the shaft and assembly are not constra<strong>in</strong>ed to rotate at a constant angular velocity 0 ,that<br />

is, it is allowed to rotate freely at angular velocity ˙. What is the difference <strong>in</strong> the overall motion?<br />

13. A rigid straight, frictionless, massless, rod rotates about the axis at an angular velocity ˙. Amass slides<br />

along the frictionless rod and is attached to the rod by a massless spr<strong>in</strong>g of spr<strong>in</strong>g constant .<br />

(a) Derive the Lagrangian and the Hamiltonian<br />

(b) Derive the equations of motion <strong>in</strong> the stationary frame us<strong>in</strong>g Hamiltonian mechanics.<br />

(c) What are the constants of motion?<br />

(d) If the rotation is constra<strong>in</strong>ed to have a constant angular velocity ˙ = then is the non-cyclic Routhian<br />

= − ˙ a constant of motion, and does it equal the total energy?<br />

(e) Use the non-cyclic Routhian toderivetheradialequationofmotion<strong>in</strong>therotat<strong>in</strong>gframeof<br />

reference for the cranked system with ˙ = .


220 CHAPTER 8. HAMILTONIAN MECHANICS<br />

14. A th<strong>in</strong> uniform rod of length 2 and mass is suspended from a massless str<strong>in</strong>g of length tiedtoanail.<br />

Initially the rod hangs vertically. A weak horizontal force is applied to the rod’s free end.<br />

(a) Write the Lagrangian for this system.<br />

(b) For very short times such that all angles are small, determ<strong>in</strong>e the angles that str<strong>in</strong>g and the rod make<br />

with the vertical. Start from rest at =0<br />

(c) Draw a diagram to illustrate the <strong>in</strong>itial motion of the rod.<br />

15. A uniform ladder of mass and length 2 is lean<strong>in</strong>g aga<strong>in</strong>st a frictionless vertical wall with its feet on a<br />

frictionless horizontal floor. Initially the stationary ladder is released at an angle 0 =60 ◦ to the floor. Assume<br />

that gravitation field =981 2 acts vertically downward and that the moment of <strong>in</strong>ertia of the ladder<br />

about its midpo<strong>in</strong>t is = 1 3 2 .<br />

(a) What is the number of generalized coord<strong>in</strong>ates necessary to describe the problem?<br />

(b) Derive the Lagrangian<br />

(c) Derive the Hamiltonian<br />

(d) Expla<strong>in</strong> if the Hamiltonian is conserved and/or if it equals the total energy<br />

(e) Use the Lagrangian to derive the equations of motion<br />

(f) Derive the angle at which the ladder loses contact with the vertical wall?<br />

16. The classical mechanics exam <strong>in</strong>duces Jacob to try his hand at bungee jump<strong>in</strong>g. Assume Jacob’s mass is<br />

suspended <strong>in</strong> a gravitational field by the bungee of unstretched length and spr<strong>in</strong>g constant . Besides the<br />

longitud<strong>in</strong>al oscillations due to the bungee jump, Jacob also sw<strong>in</strong>gs with plane pendulum motion <strong>in</strong> a vertical<br />

plane. Use polar coord<strong>in</strong>ates , neglect air drag, and assume that the bungee always is under tension.<br />

(a) Derive the Lagrangian<br />

(b) Determ<strong>in</strong>e Lagrange’s equation of motion for angular motion and identify by name the forces contribut<strong>in</strong>g<br />

to the angular motion.<br />

(c) Determ<strong>in</strong>e Lagrange’s equation of motion for radial oscillation and identify by name the forces contribut<strong>in</strong>g<br />

to the tension <strong>in</strong> the spr<strong>in</strong>g.<br />

(d) Derive the generalized momenta<br />

(e) Determ<strong>in</strong>e the Hamiltonian and give all of Hamilton’s equations of motion.


Chapter 9<br />

Hamilton’s Action Pr<strong>in</strong>ciple<br />

9.1 Introduction<br />

Hamilton’s pr<strong>in</strong>ciple of stationary action was <strong>in</strong>troduced <strong>in</strong> two papers published by Hamilton <strong>in</strong> 1834 and<br />

1835 As mentioned <strong>in</strong> the Prologue, Hamilton’s Action Pr<strong>in</strong>ciple is the foundation of the hierarchy of three<br />

philosophical stages that are used <strong>in</strong> apply<strong>in</strong>g analytical mechanics. The first stage is to use Hamilton’s<br />

Action Pr<strong>in</strong>ciple to derive either the Hamiltonian and Lagrangian for the system. The second stage is to use<br />

either Lagrangian mechanics, or Hamiltonian mechanics, to derive the equations of motion for the system.<br />

The third stage is to solve these equations of motion for the assumed <strong>in</strong>itial conditions. Lagrange had<br />

pioneered Lagrangian mechanics <strong>in</strong> 1788 based on d’Alembert’s Pr<strong>in</strong>ciple. Hamilton’s Action Pr<strong>in</strong>ciple now<br />

underlies theoretical physics, and many other discipl<strong>in</strong>es <strong>in</strong> mathematics and economics. In 1834 Hamilton<br />

was seek<strong>in</strong>g a theory of optics when he developed both his Action Pr<strong>in</strong>ciple, and the field of Hamiltonian<br />

mechanics.<br />

Hamilton’s Action Pr<strong>in</strong>ciple is based on def<strong>in</strong><strong>in</strong>g the action functional 1 for generalized coord<strong>in</strong>ates<br />

which are expressed by the vector q and their correspond<strong>in</strong>g velocity vector ˙q.<br />

=<br />

Z <br />

<br />

(q ˙q) (9.1)<br />

The scalar action is a functional of the Lagrangian (q ˙q), <strong>in</strong>tegrated between an <strong>in</strong>itial time and<br />

f<strong>in</strong>al time . In pr<strong>in</strong>ciple, higher order time derivatives of the generalized coord<strong>in</strong>ates could be <strong>in</strong>cluded, but<br />

most systems <strong>in</strong> classical mechanics are described adequately by <strong>in</strong>clud<strong>in</strong>g only the generalized coord<strong>in</strong>ates,<br />

plus their velocities. The def<strong>in</strong>ition of the action functional allows for more general Lagrangians than the<br />

simple Standard Lagrangian (q ˙q) = ( ˙q) − (q) that has been used throughout chapters 5 − 8.<br />

Hamilton stated that the actual trajectory of a mechanical system is that given by requir<strong>in</strong>g that the action<br />

functional is stationary with respect to change of the variables. The action functional is stationary when the<br />

variational pr<strong>in</strong>ciple can be written <strong>in</strong> terms of a virtual <strong>in</strong>f<strong>in</strong>itessimal displacement, to be<br />

Z <br />

= (q ˙q) =0 (9.2)<br />

<br />

Typically the stationary po<strong>in</strong>t corresponds to a m<strong>in</strong>imum of the action functional. Apply<strong>in</strong>g variational<br />

calculus to the action functional leads to the same Lagrange equations of motion for systems as the equations<br />

derived us<strong>in</strong>g d’Alembert’s Pr<strong>in</strong>ciple, if the additional generalized force terms, P <br />

=1 <br />

<br />

(q)+ <br />

,<br />

are omitted <strong>in</strong> the correspond<strong>in</strong>g equations of motion.<br />

These are used to derive the equations of motion, which then are solved for an assumed set of <strong>in</strong>itial<br />

conditions. Prior to Hamilton’s Action Pr<strong>in</strong>ciple, Lagrange developed Lagrangian mechanics based on<br />

d’Alembert’s Pr<strong>in</strong>ciple while the Newtonian equations of motion are def<strong>in</strong>ed <strong>in</strong> terms of Newton’s Laws of<br />

Motion.<br />

1 The term "action functional" was named "Hamilton’s Pr<strong>in</strong>cipal Function" <strong>in</strong> older texts. The name usually is abbreviated<br />

to "action" <strong>in</strong> modern mechanics.<br />

221


222 CHAPTER 9. HAMILTON’S ACTION PRINCIPLE<br />

9.2 Hamilton’s Pr<strong>in</strong>ciple of Stationary Action<br />

Hamilton’s crown<strong>in</strong>g achievement was his use of the general form of<br />

Hamilton’s pr<strong>in</strong>ciple of stationary action , equation92, toderive<br />

both Lagrangian mechanics, and Hamiltonian mechanics. Consider<br />

the action for the extremum path of a system <strong>in</strong> configuration<br />

space, that is, along path for =1 2 coord<strong>in</strong>ates ( ) at<br />

<strong>in</strong>itial time to ( ) at a f<strong>in</strong>al time as shown <strong>in</strong> figure 91.<br />

Then the action is given by<br />

Z <br />

= (q() ˙q()) (9.3)<br />

<br />

As used <strong>in</strong> chapter 52 a family of neighbor<strong>in</strong>g paths is def<strong>in</strong>ed<br />

by add<strong>in</strong>g an <strong>in</strong>f<strong>in</strong>itessimal fraction of a cont<strong>in</strong>uous, well-behaved<br />

neighbor<strong>in</strong>g function where =0for the extremum path. That<br />

is,<br />

( ) = ( 0) + () (9.4)<br />

In contrast to the variational case discussed when deriv<strong>in</strong>g Lagrangian<br />

mechanics, the variational path used here does not assume<br />

that the functions () vanish at the end po<strong>in</strong>ts. Assume that the<br />

neighbor<strong>in</strong>g path has an action where<br />

=<br />

Z +∆<br />

+∆<br />

(q()+δq() ˙q()+δ ˙q()) (9.5)<br />

q<br />

j<br />

q<br />

k<br />

t = t 2<br />

q (t)<br />

j<br />

t = t 2<br />

A<br />

q (t) q (t)<br />

j j<br />

B<br />

t = t 1 t = t 1 t1<br />

Figure 9.1: Extremum path A, plus<br />

the neighbor<strong>in</strong>g path B, shown <strong>in</strong> configuration<br />

space.<br />

Expand<strong>in</strong>g the <strong>in</strong>tegrand of <strong>in</strong> equation 95 gives that, relative to the extremum path , the <strong>in</strong>cremental<br />

change <strong>in</strong> action is<br />

Z X<br />

µ <br />

= − =<br />

+ <br />

˙ +[∆] <br />

˙ <br />

(9.6)<br />

<br />

The second term <strong>in</strong> the <strong>in</strong>tegral can be <strong>in</strong>tegrated by parts s<strong>in</strong>ce ˙ = <br />

=<br />

Z <br />

<br />

<br />

<br />

⎡<br />

X<br />

µ <br />

− <br />

<br />

+ ⎣ X<br />

<br />

˙ <br />

<br />

³ ´<br />

<br />

<br />

lead<strong>in</strong>g to<br />

⎤<br />

<br />

+ ∆⎦<br />

˙ <br />

<br />

<br />

(9.7)<br />

Note that equation 97 <strong>in</strong>cludes contributions from the entire path of the <strong>in</strong>tegral as well as the variations<br />

at the ends of the curve and the ∆ terms. Equation 97 leads to the follow<strong>in</strong>g two pioneer<strong>in</strong>g pr<strong>in</strong>ciples of<br />

least action <strong>in</strong> variational mechanics that were developed by Hamilton.<br />

9.2.1 Stationary-action pr<strong>in</strong>ciple <strong>in</strong> Lagrangian mechanics<br />

Derivation of Lagrangian mechanics <strong>in</strong> chapter 6 was based on the extremum path for neighbor<strong>in</strong>g paths<br />

between two given locations q( ) and q( ) that the system occupies at the <strong>in</strong>itial and f<strong>in</strong>al times and <br />

respectively. For the special case, where the end po<strong>in</strong>ts do not vary, that is, when ( )= ( )=0,and<br />

∆ = ∆ =0, then the least action for the stationary path (98) reduces to<br />

Z X<br />

µ <br />

=<br />

− <br />

<br />

=0 (9.8)<br />

˙ <br />

<br />

<br />

For <strong>in</strong>dependent generalized coord<strong>in</strong>ates , the <strong>in</strong>tegrand <strong>in</strong> brackets vanishes lead<strong>in</strong>g to the Euler-Lagrange<br />

equations. Conversely, if the Euler-Lagrange equations <strong>in</strong> 98 are satisfied, then, =0 that is, the path<br />

is stationary. This leads to the statement that the path <strong>in</strong> configuration space between two configurations<br />

q( ) and q( ) that the system occupies at times and respectively, is that for which the action is<br />

stationary. This is a statement of Hamilton’s Pr<strong>in</strong>ciple.<br />

t2<br />

q<br />

i


9.2. HAMILTON’S PRINCIPLE OF STATIONARY ACTION<br />

9.2.2 Stationary-action pr<strong>in</strong>ciple <strong>in</strong> Hamiltonian mechanics<br />

Hamilton used the general variation of the least-action path to derive of the basic equations of Hamiltonian<br />

mechanics. For the general path, the <strong>in</strong>tegral term <strong>in</strong> equation 97 vanishes because the Euler-Lagrange<br />

equations are obeyed for the stationary path. Thus the only rema<strong>in</strong><strong>in</strong>g non-zero contributions are due to<br />

the end po<strong>in</strong>t terms, which can be written by def<strong>in</strong><strong>in</strong>g the total variation of each end po<strong>in</strong>t to be<br />

where and ˙ are evaluated at and .Thenequation97 reduces to<br />

⎡<br />

= ⎣ X <br />

⎤<br />

<br />

+ ∆⎦<br />

˙ <br />

<br />

∆ = + ˙ ∆ (9.9)<br />

<br />

=<br />

⎡<br />

⎣ X <br />

⎛<br />

<br />

∆ + ⎝− X ˙ <br />

<br />

⎞ ⎤<br />

<br />

˙ + ⎠ ∆⎦<br />

˙ <br />

<br />

<br />

(9.10)<br />

S<strong>in</strong>ce the generalized momentum = <br />

˙ <br />

,thenequation910 can be expressed <strong>in</strong> terms of the Hamiltonian<br />

and generalized momentum as<br />

⎡<br />

⎤<br />

=<br />

⎣ X <br />

∆ − ∆⎦<br />

<br />

<br />

=[p·∆q − ∆] <br />

<br />

(9.11)<br />

<br />

= = (9.12)<br />

˙ <br />

Equation 911 conta<strong>in</strong>s Hamilton’s Pr<strong>in</strong>ciple of Least-action. Equation 912 gives an alternative relation of<br />

the generalized momentum that is expressed <strong>in</strong> terms of the action functional . Note that equations<br />

911 and 912 were derived directly without <strong>in</strong>vok<strong>in</strong>g reference to the Lagrangian.<br />

Integrat<strong>in</strong>g the action , equation910, between the end po<strong>in</strong>ts gives the action for the path between<br />

= and = ,thatis,( ( ) 1 ( ) 2 ) to be<br />

( ( ) ( ) )=<br />

The stationary path is obta<strong>in</strong>ed by us<strong>in</strong>g the variational pr<strong>in</strong>ciple<br />

= <br />

Z <br />

<br />

Z <br />

<br />

[p · ˙q − (q p)] (9.13)<br />

[p · ˙q − (q p)] =0 (9.14)<br />

The <strong>in</strong>tegrand, =[p · ˙q − (q p)] <strong>in</strong> this modified Hamilton’s pr<strong>in</strong>ciple, can be used <strong>in</strong> the Euler-<br />

Lagrange equations for =1 2 3 to give<br />

µ <br />

<br />

− = ˙ + =0 (9.15)<br />

˙ <br />

Similarly, the other Euler-Lagrange equations give<br />

<br />

<br />

µ <br />

˙ <br />

<br />

− <br />

<br />

= − ˙ + <br />

<br />

=0 (9.16)<br />

Thus Hamilton’s pr<strong>in</strong>ciple of least-action leads to Hamilton’s equations of motion, that is equations 915<br />

and 916.<br />

The total time derivative of the action , which is a function of the coord<strong>in</strong>ates and time, is<br />

<br />

= <br />

+ X <br />

˙ = <br />

+ p · ˙q (9.17)<br />

But the total time derivative of equation 914 equals<br />

<br />

<br />

<br />

= p · ˙q − (q p) (9.18)


224 CHAPTER 9. HAMILTON’S ACTION PRINCIPLE<br />

Comb<strong>in</strong><strong>in</strong>g equations 917 and 918 gives the Hamilton-Jacobi equation which is discussed <strong>in</strong> chapter 154.<br />

<br />

+ (q p) =0 (9.19)<br />

<br />

In summary, Hamilton’s pr<strong>in</strong>ciple of least action leads directly to Hamilton’s equations of motion (915 916)<br />

plus the Hamilton-Jacobi equation (919). Note that the above discussion has derived both Hamilton’s Action<br />

Pr<strong>in</strong>ciple (98) and Hamilton’s equations of motion (915 916) directly from Hamilton’s variational<br />

concept of stationary action, , without explicitly <strong>in</strong>vok<strong>in</strong>g the Lagrangian.<br />

9.2.3 Abbreviated action<br />

Hamilton’s Action Pr<strong>in</strong>ciple determ<strong>in</strong>es completely the path of the motion and the position on the path as<br />

a function of time. If the Lagrangian and the Hamiltonian are time <strong>in</strong>dependent, that is, conservative, then<br />

= and equation 913 equals<br />

The R 2<br />

1<br />

( ( 1 ) 1 ( 2 ) 2 )=<br />

Z <br />

<br />

[p · ˙q − ] =<br />

Z <br />

<br />

p·q − ( − ) (9.20)<br />

p · δ ˙q term <strong>in</strong> equation 920, is called the abbreviated action which is def<strong>in</strong>ed as<br />

0 ≡<br />

Z <br />

<br />

p· ˙q =<br />

Z <br />

<br />

p·q (9.21)<br />

The abbreviated action can be simplified assum<strong>in</strong>g use of the standard Lagrangian = − with a<br />

velocity-<strong>in</strong>dependent potential , thenequation84 gives.<br />

Z <br />

X<br />

Z <br />

0 ≡ ˙ = ( + ) = 2 = p·q (9.22)<br />

<br />

<br />

Abbreviated action provides for use of a simplified form of the pr<strong>in</strong>ciple of least action that is based<br />

on the k<strong>in</strong>etic energy, and not potential energy. For conservative systems it determ<strong>in</strong>es the path of the<br />

motion, but not the time dependence of the motion. Consider virtual motions where the path satisfies<br />

energy conservation, and where the end po<strong>in</strong>ts are held fixed, that is =0 but allow for a variation <strong>in</strong><br />

the f<strong>in</strong>al time. Then us<strong>in</strong>g the Hamilton-Jacobi equation, 919<br />

Z <br />

Z <br />

= − = − (9.23)<br />

However, equation 921 gives that<br />

= 0 − (9.24)<br />

Therefore<br />

0 =0 (9.25)<br />

That is, the abbreviated action has a m<strong>in</strong>imum with respect to all paths that satisfy the conservation of<br />

energy which can be written as<br />

Z <br />

0 = 2 =0 (9.26)<br />

<br />

Equation 926 is called the Maupertuis’ least-action pr<strong>in</strong>ciple whichheproposed<strong>in</strong>1744 based on Fermat’s<br />

Pr<strong>in</strong>ciple <strong>in</strong> optics. Credit for the formulation of least action commonly is given to Maupertuis; however, the<br />

Maupertuis pr<strong>in</strong>ciple is similar to the use of least action applied to the “vis viva”, as was proposed by Leibniz<br />

four decades earlier. Maupertuis used teleological arguments, rather than scientific rigor, because of his<br />

limited mathematical capabilities. In 1744 Euler provided a scientifically rigorous argument, presented above,<br />

that underlies the Maupertuis pr<strong>in</strong>ciple. Euler derived the correct variational relation for the abbreviated<br />

action to be<br />

Z X <br />

0 = =0 (9.27)<br />

<br />

Hamilton’s use of the pr<strong>in</strong>ciple of least action to derive both Lagrangian and Hamiltonian mechanics is<br />

a remarkable accomplishment. It underlies both Lagrangian and Hamiltonian mechanics and confirmed the<br />

conjecture of Maupertuis.


9.2. HAMILTON’S PRINCIPLE OF STATIONARY ACTION<br />

9.2.4 Hamilton’s Pr<strong>in</strong>ciple applied us<strong>in</strong>g <strong>in</strong>itial boundary conditions<br />

Galley[Gal13] identified a subtle <strong>in</strong>consistency <strong>in</strong> the applications<br />

of Hamilton’s Pr<strong>in</strong>ciple of Stationary Action to both<br />

Lagrangian and Hamiltonian mechanics. The <strong>in</strong>consistency<br />

<strong>in</strong>volves the fact that Hamilton’s Pr<strong>in</strong>ciple is def<strong>in</strong>ed as the<br />

action <strong>in</strong>tegral between the <strong>in</strong>itial time and the f<strong>in</strong>al time<br />

as boundary conditions, that is, it is assumed to be time<br />

symmetric. However, most applications <strong>in</strong> Lagrangian and<br />

Hamiltonian mechanics assume that the action <strong>in</strong>tegral is<br />

evaluated based on the <strong>in</strong>itial values as the boundary conditions,<br />

rather than the <strong>in</strong>itial and f<strong>in</strong>al times . That<br />

is, typical applications require use of a time-asymmetric<br />

version of Hamilton’s pr<strong>in</strong>ciple. Galley[Gal13][Gal14] proposed<br />

a framework for transform<strong>in</strong>g Hamilton’s Pr<strong>in</strong>ciple to<br />

a time-asymmetric form <strong>in</strong> order to handle problems where<br />

the boundary conditions are based on us<strong>in</strong>g only the <strong>in</strong>itial<br />

values at the <strong>in</strong>itial time , rather than the <strong>in</strong>itial plus<br />

f<strong>in</strong>al times ( ) that is assumed <strong>in</strong> the time-symmetric<br />

def<strong>in</strong>ition of the action <strong>in</strong> Hamilton’s Pr<strong>in</strong>ciple.<br />

The follow<strong>in</strong>g describes the framework proposed by<br />

Galley for transform<strong>in</strong>g Hamilton’s Pr<strong>in</strong>ciple to a timeasymmetric<br />

form. Let q and ˙q designate sets of generalized<br />

coord<strong>in</strong>ates, plus their velocities, where q and ˙q are<br />

the fundamental variables assumed <strong>in</strong> the def<strong>in</strong>ition of the<br />

Figure 9.2: The left schematic shows paths between<br />

the <strong>in</strong>itial q( ) and f<strong>in</strong>al q( ) times<br />

for conservative mechanics. The solid l<strong>in</strong>e designates<br />

the path for which the action is stationary,<br />

while the dashed l<strong>in</strong>es represent the<br />

varied paths. The right schematic shows the<br />

paths applied to the doubled degrees of freedom<br />

with two <strong>in</strong>itial boundary conditions, that<br />

is, q 1 ( ) and q 2 ( ) plus assum<strong>in</strong>g that both<br />

paths are identical at their <strong>in</strong>tersection and<br />

that they <strong>in</strong>tersect at the same f<strong>in</strong>al time, that<br />

is, q 1 ( )=q 2 ( ).<br />

Lagrangian used by Hamilton’s Pr<strong>in</strong>ciple. As illustrated<br />

schematically <strong>in</strong> figure 92, Galley proposed doubl<strong>in</strong>g the<br />

number of degrees of freedom for the system considered, that is, let q → (q 1 q 2 ) and ˙q → ( ˙q 1 ˙q 2 ). In addition<br />

he def<strong>in</strong>es two identical variational paths 1 and 2 where path 2 is the time reverse of path 1. That<br />

is, path 1 starts at the <strong>in</strong>itial time ,andendsat , whereas path 2 starts at and ends at . That<br />

is, he assumes that q and ˙q specify the two paths <strong>in</strong> the space of the doubled degrees of freedom that are<br />

identical, and that they <strong>in</strong>tersect at the f<strong>in</strong>al time . The arrows shown on the paths <strong>in</strong> figure 92 designate<br />

the assumed direction of the time <strong>in</strong>tegration along these paths.<br />

For the doubled system of degrees of freedom, the total action for the sum of the two paths is given by<br />

the time <strong>in</strong>tegral of the doubled variables, (q 1 q 2 ) which can be written as<br />

(q 1 q 2 )=<br />

Z <br />

<br />

(q 1 ˙q 1 ) +<br />

Z <br />

<br />

(q 2 ˙q 2 ) =<br />

Z <br />

<br />

[ (q 1 ˙q 1 ) − (q 2 ˙q 2 )] (9.28)<br />

The above relation assumes that the doubled variables (q 1 ˙q 1 ) and (q 2 ˙q 2 ) are decoupled from each other.<br />

More generally one can assume that the two sets of variables are coupled by some arbitrary function<br />

(q 1 ˙q 1 q 2 ˙q 2 ). Then the action can be written as<br />

Z <br />

(q 1 q 2 )= [ (q 1 ˙q 1 t) − (q 2 ˙q 2 t)+ (q 1 ˙q 1 q 2 ˙q 2 )] (9.29)<br />

<br />

The effective Lagrangian for this doubled system then can be def<strong>in</strong>ed as<br />

and the action can be written as<br />

Λ (q 1 q 2 ˙q 1 ˙q 2 ) ≡ [ (q 1 ˙q 1 ) − (q 2 ˙q 2 )+ (q 1 ˙q 1 q 2 ˙q 2 )] (9.30)<br />

(q 1 q 2 )=<br />

Z <br />

<br />

Λ (q 1 ˙q 1 q 2 ˙q 2 ) (9.31)<br />

The coupl<strong>in</strong>g term (q 1 ˙q 1 q 2 ˙q 2 ) for the doubled system of degrees of freedom must satisfy the<br />

follow<strong>in</strong>g two properties.


226 CHAPTER 9. HAMILTON’S ACTION PRINCIPLE<br />

(a) If it can be expressed as the difference of two scalar potentials, ∆ (q 1 q 2 )= (q 1 ) − (q 2 ),then<br />

it can be absorbed <strong>in</strong>to the potential term for each of the doubled variables <strong>in</strong> the Lagrangian. This implies<br />

that =0 and there is no reason to double the number of degrees of freedom because the system is<br />

conservative. Thus describes generalized forces that are not derivable from potential energy, that is, not<br />

conservative.<br />

(b) A second property of the coupl<strong>in</strong>g term (q 1 ˙q 1 q 2 ˙q 2 ) is that it must be antisymmetric under<br />

<strong>in</strong>terchange of the arbitrary labels 1 ↔ 2. Thatis,<br />

(q 2 ˙q 2 q 1 ˙q 1 )=− (q 1 ˙q 1 q 2 ˙q 2 ) (9.32)<br />

Therefore the antisymmetric function (q 1 ˙q 1 q 2 ˙q 2 ) vanishes when q 2 = q 1 .<br />

The variational condition requires that the action (q 1 q 2 ) has a well def<strong>in</strong>ed stationary po<strong>in</strong>t for the<br />

doubled system. This is achieved by parametriz<strong>in</strong>g both coord<strong>in</strong>ate paths as<br />

q 12 ( ) =q 12 ( 0) + 12 () (9.33)<br />

where q 12 ( 0) are the coord<strong>in</strong>ates for which the action is stationary, ¿ 1 and where 12 () are arbitrary<br />

functions of time denot<strong>in</strong>g virtual displacements of the paths. The doubled system has two <strong>in</strong>dependent<br />

paths connect<strong>in</strong>g the two <strong>in</strong>itial boundary conditions at , and it requires that these paths <strong>in</strong>tersect at .<br />

The variational system for the two <strong>in</strong>tersect<strong>in</strong>g paths requires specify<strong>in</strong>g four conditions, two per path. Two<br />

of the four conditions are determ<strong>in</strong>ed by requir<strong>in</strong>g that at the <strong>in</strong>itial boundary conditions satisfies that<br />

12 ( )=0. The rema<strong>in</strong><strong>in</strong>g two conditions are derived by requir<strong>in</strong>g that the variation of the action (q 1 q 2 )<br />

satisfies<br />

∙ ¸ Z <br />

<br />

½ ∙ Λ<br />

=0= <br />

<br />

1 − ¸ ∙<br />

1<br />

Λ<br />

− <br />

=0<br />

<br />

1 <br />

2 − ¸ ¾<br />

2<br />

+[<br />

=0<br />

2 <br />

1 1 − 2 2 ] =<br />

(9.34)<br />

=0<br />

The canonical momenta 12 conjugate to the doubled coord<strong>in</strong>ates q 12 are def<strong>in</strong>ed us<strong>in</strong>g the nonconservative<br />

Lagrangian Λ to be<br />

1 (q 12 ˙q 12 ) ≡<br />

Λ<br />

˙ 1 () = (q 1 ˙q 1 )<br />

˙ 1 ()<br />

+ (q 1 ˙q 1 q 2 ˙q 2 )<br />

˙ 1 () (9.35)<br />

where the superscript designates the solution based on the <strong>in</strong>itial conditions. Note that the conjugate<br />

momentum 1 = (q 1 ˙q 1 )<br />

˙ 1 ()<br />

while the (q 1 ˙q 1 q 2 ˙q 2 )<br />

˙ 1 ()<br />

term is part of the total momentum due to the<br />

nonconservative <strong>in</strong>teraction. Similarly the momentum for the second path is<br />

2 (q 12 ˙q 12 ) ≡<br />

Λ<br />

˙ 2 () = (q 1 ˙q 1 )<br />

˙ 2 ()<br />

+ (q 1 ˙q 1 q 2 ˙q 2 )<br />

˙ 2 () (9.36)<br />

The last term <strong>in</strong> equation 934 that is, the term [ 1 1 − 2 2 ] =<br />

results from <strong>in</strong>tegration by parts,<br />

which will vanish if<br />

1( ) 1( )= 2( ) 2( ) (9.37)<br />

The equality condition at the <strong>in</strong>tersection of the two paths at requires that<br />

Therefore equations 937 and 938 imply that<br />

1( )= 2( ) (9.38)<br />

1( )= 2( ) (9.39)<br />

Therefore equations 938 and 939 constitute the equality condition that must be satisfied when the two<br />

paths <strong>in</strong>tersect at . The equality condition ensures that the boundary term for <strong>in</strong>tegration by parts <strong>in</strong><br />

equation 934 will vanish for arbitrary variations provided that the two unspecified paths agree at the f<strong>in</strong>al<br />

time . Similarly the conjugate momenta 1( ) 2( ) must agree, but otherwise are unspecified. As a<br />

consequence, the equality condition ensures that the variational pr<strong>in</strong>ciple is consistent with the f<strong>in</strong>al state at


9.2. HAMILTON’S PRINCIPLE OF STATIONARY ACTION<br />

not be<strong>in</strong>g specified. That is, the equations of motion are only specified by the <strong>in</strong>itial boundary conditions<br />

ofthetimeasymmetricactionforthedoubledsystem.<br />

More physics <strong>in</strong>sight is provided by us<strong>in</strong>g a more convenient parametrization of the coord<strong>in</strong>ates <strong>in</strong> terms<br />

of their average and difference. That is, let<br />

+ ≡ 1 + 2<br />

− ≡ 1 − 2 (9.40)<br />

2<br />

Then the physical limit is<br />

+ → − → 0 (9.41)<br />

That is, the average history is the relevant physical history, while the difference coord<strong>in</strong>ate simply vanishes.<br />

For these coord<strong>in</strong>ates, the nonconservative Lagrangian is Λ (q + q − ˙q + ˙q − ) and the equality conditions<br />

reduce to<br />

− ( ) = 0 (9.42)<br />

− ( ) = 0 (9.43)<br />

which implies that the physically relevant average (+) quantities are not specified at the f<strong>in</strong>al time <br />

order to have a well-def<strong>in</strong>ed variational pr<strong>in</strong>ciple.<br />

The canonical momenta are given by<br />

+ = 1 + 2<br />

2<br />

= Λ<br />

˙ −<br />

<strong>in</strong><br />

(9.44)<br />

− = 1 − 2 = Λ<br />

˙ +<br />

(9.45)<br />

The equations of motion can be written as<br />

Λ<br />

˙ ±<br />

= Λ<br />

±<br />

(9.46)<br />

Equation 946 is identically zero for the + subscript, while, <strong>in</strong> the physical limit (PL), the negative subscript<br />

gives that<br />

∙ Λ<br />

˙ −<br />

− Λ ¸<br />

−<br />

=0<br />

<br />

(9.47)<br />

Substitut<strong>in</strong>g for the Lagrangian Λ gives that<br />

<br />

˙ −<br />

− ∙ <br />

−<br />

=<br />

− − ¸<br />

<br />

˙ −<br />

≡ (q 1 ˙q 1 )<br />

<br />

(9.48)<br />

where is a generalized nonconservative force derived from .<br />

Note that equation 946 can be derived equally well by tak<strong>in</strong>g the direct functional derivative with respect<br />

to −(), thatis,<br />

∙ <br />

0=<br />

− ()¸<br />

(9.49)<br />

The above time-asymmetric formalism applies Hamilton’s action pr<strong>in</strong>ciple to systems that <strong>in</strong>volve <strong>in</strong>itial<br />

boundary conditions while the second path corresponds to the f<strong>in</strong>al boundary conditions. This framework,<br />

proposed recently by Galley[Gal13], provides a remarkable advance for the handl<strong>in</strong>g of nonconservative action<br />

<strong>in</strong> Lagrangian and Hamiltonian mechanics. 2 This formalism directly <strong>in</strong>corporates the variational pr<strong>in</strong>ciple<br />

for <strong>in</strong>itial boundary conditions and causal dynamics that are usually required for applications of Lagrangian<br />

and Hamiltonian mechanics. Currently, there is limited exploitation of this new formalism because there<br />

has been <strong>in</strong>sufficient time for it to become well known, for full recognition of its importance, and for the<br />

development and publication of applications. Chapter 10 discusses an application of this formalism to<br />

nonconservative systems <strong>in</strong> classical mechanics.<br />

2 This topic goes beyond the planned scope of this book. It is recommended that the reader refer to the work of Galley,<br />

Tsang, and Ste<strong>in</strong>[Gal13, Gal14] for further discussion plus examples of apply<strong>in</strong>g this formalism to nonconservative systems <strong>in</strong><br />

classical mechanics, electromagnetic radiation, RLC circuits, fluid dynamics, and field theory.


228 CHAPTER 9. HAMILTON’S ACTION PRINCIPLE<br />

9.3 Lagrangian<br />

9.3.1 Standard Lagrangian<br />

Lagrangian mechanics, as <strong>in</strong>troduced <strong>in</strong> chapter 6 was based on the concepts of k<strong>in</strong>etic energy and potential<br />

energy. d’Alembert’s pr<strong>in</strong>ciple of virtual work was used to derive Lagrangian mechanics <strong>in</strong> chapter 6 and this<br />

led to the def<strong>in</strong>ition of the standard Lagrangian. That is, the standard Lagrangian was def<strong>in</strong>ed <strong>in</strong> chapter<br />

62 to be the difference between the k<strong>in</strong>etic and potential energies.<br />

(q ˙q) = ( ˙q) − (q) (9.50)<br />

Hamilton extended Lagrangian mechanics by def<strong>in</strong><strong>in</strong>g Hamilton’s Pr<strong>in</strong>ciple, equation 92, which states that<br />

a dynamical system follows a path for which the action functional is stationary, that is, the time <strong>in</strong>tegral<br />

of the Lagrangian. Chapter 6 showed that us<strong>in</strong>g the standard Lagrangian for def<strong>in</strong><strong>in</strong>g the action functional<br />

leads to the Euler-Lagrange variational equations<br />

½ <br />

<br />

µ <br />

− ¾<br />

= <br />

+<br />

˙ <br />

X<br />

=1<br />

<br />

<br />

<br />

(q) (9.51)<br />

The Lagrange multiplier terms handle the holonomic constra<strong>in</strong>t forces and <br />

handles the rema<strong>in</strong><strong>in</strong>g<br />

excluded generalized forces. Chapters 6 − 8 showed that the use of the standard Lagrangian, with the Euler-<br />

Lagrange equations (951) provides a remarkably powerful and flexible way to derive second-order equations<br />

of motion for dynamical systems <strong>in</strong> classical mechanics.<br />

Note that the Euler-Lagrange equations, expressed solely <strong>in</strong> terms of the standard Lagrangian (951)<br />

that is, exclud<strong>in</strong>g the <br />

<br />

+ P <br />

=1 <br />

<br />

(q) terms, are valid only under the follow<strong>in</strong>g conditions:<br />

1. The forces act<strong>in</strong>g on the system, apart from any forces of constra<strong>in</strong>t, must be derivable from scalar<br />

potentials.<br />

2. The equations of constra<strong>in</strong>t must be relations that connect the coord<strong>in</strong>ates of the particles and may<br />

be functions of time, that is, the constra<strong>in</strong>ts are holonomic.<br />

The <br />

<br />

+ P <br />

=1 <br />

<br />

(q) terms extend the range of validity of us<strong>in</strong>g the standard Lagrangian <strong>in</strong> the<br />

Lagrange-Euler equations by <strong>in</strong>troduc<strong>in</strong>g constra<strong>in</strong>t and omitted forces explicitly.<br />

Chapters 6−8 exploited Lagrangian mechanics based on use of the standard def<strong>in</strong>ition of the Lagrangian.<br />

The present chapter will show that the powerful Lagrangian formulation, us<strong>in</strong>g the standard Lagrangian,<br />

can be extended to <strong>in</strong>clude alternative non-standard Lagrangians that may be applied to dynamical systems<br />

whereuseofthestandarddef<strong>in</strong>ition of the Lagrangian is <strong>in</strong>applicable. If these non-standard Lagrangians<br />

satisfy Hamilton’s Action Pr<strong>in</strong>ciple, 92, then they can be used with the Euler-Lagrange equations to generate<br />

the correct equations of motion, even though the Lagrangian may not have the simple relation to the k<strong>in</strong>etic<br />

and potential energies adopted by the standard Lagrangian. Currently, the development and exploitation of<br />

non-standard Lagrangians is an active field of Lagrangian mechanics.<br />

9.3.2 Gauge <strong>in</strong>variance of the standard Lagrangian<br />

Note that the standard Lagrangian is not unique <strong>in</strong> that there is a cont<strong>in</strong>uous spectrum of equivalent<br />

standard Lagrangians that all lead to identical equations of motion. This is because the Lagrangian is a<br />

scalar quantity that is <strong>in</strong>variant with respect to coord<strong>in</strong>ate transformations. The follow<strong>in</strong>g transformations<br />

change the standard Lagrangian, but leave the equations of motion unchanged.<br />

1. The Lagrangian is <strong>in</strong>def<strong>in</strong>ite with respect to addition of a constant to the scalar potential which cancels<br />

out when the derivatives <strong>in</strong> the Euler-Lagrange differential equations are applied.<br />

2. The Lagrangian is <strong>in</strong>def<strong>in</strong>ite with respect to addition of a constant k<strong>in</strong>etic energy.<br />

3. The Lagrangian is <strong>in</strong>def<strong>in</strong>ite with respect to addition of a total time derivative of the form 2 →<br />

1 + [Λ( )] for any differentiable function Λ( ) of the generalized coord<strong>in</strong>ates plus time, that<br />

has cont<strong>in</strong>uous second derivatives.


9.3. LAGRANGIAN 229<br />

This last statement can be proved by consider<strong>in</strong>g a transformation between two related standard Lagrangians<br />

of the form<br />

2 (q ) = 1 (q )+ Λ(q)<br />

µ<br />

= 1 (q Λ(q)<br />

)+ ˙ + Λ(q) <br />

(9.52)<br />

<br />

<br />

This leads to a standard Lagrangian 2 that has the same equations of motion as 1 as is shown by<br />

substitut<strong>in</strong>g equation 952 <strong>in</strong>to the Euler-Lagrange equations. That is,<br />

µ <br />

2<br />

− 2<br />

= µ 1<br />

˙ <br />

<br />

− 1<br />

+ 2 Λ(q)<br />

− 2 Λ(q)<br />

= ˙ <br />

µ 1<br />

˙ <br />

<br />

− 1<br />

<br />

(9.53)<br />

Thus even though the related Lagrangians 1 and 2 are different, they are completely equivalent <strong>in</strong> that<br />

they generate identical equations of motion.<br />

There is an unlimited range of equivalent standard Lagrangians that all lead to the same equations of<br />

motion and satisfy the requirements of the Lagrangian. That is, there is no unique choice among the wide<br />

range of equivalent standard Lagrangians expressed <strong>in</strong> terms of generalized coord<strong>in</strong>ates. This discussion is<br />

an example of gauge <strong>in</strong>variance <strong>in</strong> physics.<br />

Modern theories <strong>in</strong> physics describe reality <strong>in</strong> terms of potential fields. Gauge <strong>in</strong>variance, which also is<br />

called gauge symmetry, is a property of field theory for which different underly<strong>in</strong>g fields lead to identical<br />

observable quantities. Well-known examplesarethestaticelectricpotentialfield and the gravitational<br />

potential field where any arbitrary constant can be added to these scalar potentials with zero impact on the<br />

observed static electric field or the observed gravitational field. Gauge theories constra<strong>in</strong> the laws of physics<br />

<strong>in</strong> that the impact of gauge transformations must cancel out when expressed <strong>in</strong> terms of the observables.<br />

Gauge symmetry plays a crucial role <strong>in</strong> both classical and quantal manifestations of field theory, e.g. it is<br />

the basis of the Standard Model of electroweak and strong <strong>in</strong>teractions.<br />

Equivalent Lagrangians are a clear manifestation of gauge <strong>in</strong>variance as illustrated by equations 952 953<br />

which show that add<strong>in</strong>g any total time derivative of a scalar function Λ(q) to the Lagrangian has no<br />

observable consequences on the equations of motion. That is, although addition of the total time derivative<br />

of the scalar function Λ(q) changes the value of the Lagrangian, it does not change the equations of motion<br />

for the observables derived us<strong>in</strong>g equivalent standard Lagrangians.<br />

For Lagrangian formulations of classical mechanics, the gauge <strong>in</strong>variance is readily apparent by direct<br />

<strong>in</strong>spection of the Lagrangian.<br />

9.1 Example: Gauge <strong>in</strong>variance <strong>in</strong> electromagnetism<br />

The scalar electric potential Φ and the vector potential fields <strong>in</strong> electromagnetism are examples of gauge<strong>in</strong>variant<br />

fields. These electromagnetic-potential fields are not directly observable, that is, the electromagnetic<br />

observable quantities are the electric field and magnetic field which can be derived from the scalar and<br />

vector potential fields Φ and . An advantage of us<strong>in</strong>g the potential fields is that they reduce the problem<br />

from 6 components, 3 each for and to 4 components, one for the scalar field Φ and 3 for the vector<br />

potential . The Lagrangian for the velocity-dependent Lorentz force, given by equation 667 provides an<br />

example of gauge <strong>in</strong>variance. Equations 663 and 665 showed that the electric and magnetic fields can be<br />

expressed <strong>in</strong> terms of scalar and vector potentials Φ and A by the relations<br />

B = ∇ × A<br />

E = −∇Φ − A<br />

<br />

The equations of motion for a charge <strong>in</strong> an electromagnetic field can be obta<strong>in</strong>ed by us<strong>in</strong>g the Lagrangian<br />

= 1 v · v − (Φ − A · v)<br />

2<br />

Consider the transformations (AΦ) → (A 0 Φ 0 ) <strong>in</strong> the transformed Lagrangian 0 where<br />

A 0 = A + ∇Λ(r)<br />

Φ 0 = Φ − Λ(r)


230 CHAPTER 9. HAMILTON’S ACTION PRINCIPLE<br />

The transformed Lorentz-force Lagrangian 0 is related to the orig<strong>in</strong>al Lorentz-force Lagrangian by<br />

∙<br />

0 = + ṙ·∇Λ(r)+ Λ(r) ¸<br />

= + <br />

Λ(r)<br />

Note that the additive term Λ(r) is an exact time differential. Thus the Lagrangian 0 is gauge <strong>in</strong>variant<br />

imply<strong>in</strong>g identical equations of motion are obta<strong>in</strong>ed us<strong>in</strong>g either of these equivalent Lagrangians.<br />

The force fields E and B can be used to show that the above transformation is gauge-<strong>in</strong>variant. That is,<br />

E 0 = −∇Φ 0 − A0<br />

<br />

= −∇Φ − A<br />

= E<br />

B 0 = ∇ × A 0 = ∇ × A = B<br />

That is, the additive terms due to the scalar field Λ(r) cancel. Thus the electromagnetic force fields follow<strong>in</strong>g<br />

a gauge-<strong>in</strong>variant transformation are shown to be identical <strong>in</strong> agreement with what is <strong>in</strong>ferred directly by<br />

<strong>in</strong>spection of the Lagrangian.<br />

9.3.3 Non-standard Lagrangians<br />

The def<strong>in</strong>ition of the standard Lagrangian was based on d’Alembert’s differential variational pr<strong>in</strong>ciple. The<br />

flexibility and power of Lagrangian mechanics can be extended to a broader range of dynamical systems<br />

by employ<strong>in</strong>g an extended def<strong>in</strong>ition of the Lagrangian that is based on Hamilton’s Pr<strong>in</strong>ciple, equation 92.<br />

Note that Hamilton’s Pr<strong>in</strong>ciple was <strong>in</strong>troduced 46 years after development of the standard formulation of<br />

Lagrangian mechanics. Hamilton’s Pr<strong>in</strong>ciple provides a general def<strong>in</strong>ition of the Lagrangian that applies<br />

to standard Lagrangians, which are expressed as the difference between the k<strong>in</strong>etic and potential energies,<br />

as well as to non-standard Lagrangians where there may be no clear separation <strong>in</strong>to k<strong>in</strong>etic and potential<br />

energy terms. These non-standard Lagrangians can be used with the Euler-Lagrange equations to generate<br />

the correct equations of motion, even though they may have no relation to the k<strong>in</strong>etic and potential energies.<br />

The extended def<strong>in</strong>ition of the Lagrangian based on Hamilton’s action functional 91 can be exploited for<br />

develop<strong>in</strong>g non-standard def<strong>in</strong>itions of the Lagrangian that may be applied to dynamical systems where use<br />

of the standard def<strong>in</strong>ition is <strong>in</strong>applicable. Non-standard Lagrangians can be equally as useful as the standard<br />

Lagrangian for deriv<strong>in</strong>g equations of motion for a system. <strong>Second</strong>ly, non-standard Lagrangians, that have no<br />

energy <strong>in</strong>terpretation, are available for deriv<strong>in</strong>g the equations of motion for many nonconservative systems.<br />

Thirdly, Lagrangians are useful irrespective of how they were derived. For example, they can be used to<br />

derive conservation laws or the equations of motion. Coord<strong>in</strong>ate transformations of the Lagrangian is much<br />

simpler than that required for transform<strong>in</strong>g the equations of motion. The relativistic Lagrangian def<strong>in</strong>ed <strong>in</strong><br />

chapter 176 is a well-known example of a non-standard Lagrangian.<br />

9.3.4 Inverse variational calculus<br />

Non-standard Lagrangians and Hamiltonians are not based on the concept of k<strong>in</strong>etic and potential energies.<br />

Therefore, development of non-standard Lagrangians and Hamiltonians require an alternative approach that<br />

ensures that they satisfy Hamilton’s Pr<strong>in</strong>ciple, equation 92 which underlies the Lagrangian and Hamiltonian<br />

formulations. One useful alternative approach is to derive the Lagrangian or Hamiltonian via an<br />

<strong>in</strong>verse variational process based on the assumption that the equations of motion are known. Helmholtz developed<br />

the field of <strong>in</strong>verse variational calculus which plays an important role <strong>in</strong> development of non-standard<br />

Lagrangians. An example of this approach is use of the well-known Lorentz force as the basis for deriv<strong>in</strong>g<br />

a correspond<strong>in</strong>g Lagrangian to handle systems <strong>in</strong>volv<strong>in</strong>g electromagnetic forces. Inverse variational calculus<br />

is a branch of mathematics that is beyond the scope of this textbook. The Douglas theorem[Dou41] states<br />

that, if the three Helmholtz conditions are satisfied, then there exists a Lagrangian that, when used with the<br />

Euler-Lagrange differential equations, leads to the given set of equations of motion. Thus, it will be assumed<br />

that the <strong>in</strong>verse variational calculus technique can be used to derive a Lagrangian from known equations of<br />

motion.


9.4. APPLICATION OF HAMILTON’S ACTION PRINCIPLE TO MECHANICS 231<br />

9.4 Application of Hamilton’s Action Pr<strong>in</strong>ciple to mechanics<br />

Knowledge of the equations of motion is required to predict the response of a system to any set of <strong>in</strong>itial<br />

conditions. Hamilton’s action pr<strong>in</strong>ciple, that is built <strong>in</strong>to Lagrangian and Hamiltonian mechanics, coupled<br />

with the availability of a wide arsenal of variational pr<strong>in</strong>ciples and techniques, provides a remarkably powerful<br />

and broad approach to deriv<strong>in</strong>g the equations of motions required to determ<strong>in</strong>e the system response.<br />

As mentioned <strong>in</strong> the Prologue, derivation of the equations of motion for any system, based on Hamilton’s<br />

Action Pr<strong>in</strong>ciple, separates naturally <strong>in</strong>to a hierarchical set of three stages that differ <strong>in</strong> both sophistication<br />

and understand<strong>in</strong>g, as described below.<br />

1. Action stage: The primary “action stage” employs Hamilton’s Action functional, = R <br />

<br />

(q ˙q)<br />

to derive the Lagrangian and Hamiltonian functionals. This action stage provides the most fundamental<br />

and sophisticated level of understand<strong>in</strong>g. It <strong>in</strong>volves specify<strong>in</strong>g all the active degrees of freedom, as<br />

well as the <strong>in</strong>teractions <strong>in</strong>volved. Symmetries <strong>in</strong>corporated at this primary action stage can simplify<br />

subsequent use of the Hamiltonian and Lagrangian functionals.<br />

2. Hamiltonian/Lagrangian stage: The “Hamiltonian/Lagrangian stage” uses the Lagrangian or<br />

Hamiltonian functionals, that were derived at the action stage, <strong>in</strong> order to derive the equations of<br />

motion for the system of <strong>in</strong>terest. Symmetries, not already <strong>in</strong>corporated at the primary action stage,<br />

may be <strong>in</strong>cluded at this secondary stage.<br />

3. Equations of motion stage: The “equations-of-motion stage” uses the derived equations of motion to<br />

solve for the motion of the system subject to a given set of <strong>in</strong>itial boundary conditions. Nonconservative<br />

forces, such as dissipative forces, that were not <strong>in</strong>cluded at the primary and secondary stages, may be<br />

added at the equations of motion stage.<br />

Lagrange omitted the action stage when he used d’Alembert’s Pr<strong>in</strong>ciple to derive Lagrangian mechanics.<br />

The Newtonian mechanics approach omits both the primary “action” stage, as well as the secondary<br />

“Hamiltonian/Lagrangian” stage, s<strong>in</strong>ce Newton’s Laws of Motion directly specify the “equations-of-motion<br />

stage”. Thus these do not allow exploit<strong>in</strong>g the considerable advantages provided by the use of the action, the<br />

Lagrangian, and the Hamiltonian. Newtonian mechanics requires that all the active forces be <strong>in</strong>cluded when<br />

deriv<strong>in</strong>g the equations of motion, which <strong>in</strong>volves deal<strong>in</strong>g with vector quantities. In Newtonian mechanics,<br />

symmetries must be <strong>in</strong>corporated directly at the equations of motion stage, which is more difficult than<br />

when done at the primary “action” stage, or the secondary “Lagrangian/Hamiltonian” stage. The “action”<br />

and “Hamiltonian/Lagrangian” stages allow for use of the powerful arsenal of mathematical techniques that<br />

have been developed for apply<strong>in</strong>g variational pr<strong>in</strong>ciples.<br />

There are considerable advantages to deriv<strong>in</strong>g the equations of motion based on Hamilton’s Pr<strong>in</strong>ciple,<br />

rather than derive them us<strong>in</strong>g Newtonian mechanics. It is significantly easier to use variational pr<strong>in</strong>ciples to<br />

handle the scalar functionals, action, Lagrangian, and Hamiltonian, rather than start<strong>in</strong>g at the equationsof-motion<br />

stage. For example, utiliz<strong>in</strong>g all three stages of algebraic mechanics facilitates accommodat<strong>in</strong>g<br />

extra degrees of freedom, symmetries, and <strong>in</strong>teractions. The symmetries identified by Noether’s theorem are<br />

more easily recognized dur<strong>in</strong>g the primary “action” and secondary “Hamiltonian/Lagrangian” stages rather<br />

than at the subsequent “equations of motion” stage. Approximations made at the “action” stage are easier<br />

to implement than at the “equations-of-motion” stage. Constra<strong>in</strong>ed motion is much more easily handled at<br />

the primary “action”, or secondary “Hamilton/Lagrangian” stages, than at the equations-of-motion stage.<br />

An important advantage of us<strong>in</strong>g Hamilton’s Action Pr<strong>in</strong>ciple, is that there is a close relationship between<br />

action <strong>in</strong> classical and quantal mechanics, as discussed <strong>in</strong> chapters 15 and 18. Algebraic pr<strong>in</strong>ciples, that<br />

underly analytical mechanics, naturally encompass applications to many branches of modern physics, such<br />

as relativistic mechanics, fluid motion, and field theory.<br />

In summary, the use of the s<strong>in</strong>gle fundamental <strong>in</strong>variant quantity, action, as described above, provides a<br />

powerful and elegant framework, that was developed first for classical mechanics, but now is exploited <strong>in</strong> a<br />

wide range of science, eng<strong>in</strong>eer<strong>in</strong>g, and economics. An important feature of us<strong>in</strong>g the algebraic approach to<br />

classical mechanics is the tremendous arsenal of powerful mathematical techniques that have been developed<br />

for use of variational calculus applied to Lagrangian and Hamiltonian mechanics. Some of these variational<br />

techniques were presented <strong>in</strong> chapters 6 7 8 and 9, while others will be <strong>in</strong>troduced <strong>in</strong> chapter 15.


232 CHAPTER 9. HAMILTON’S ACTION PRINCIPLE<br />

9.5 Summary<br />

The Hamilton’s 1834 publication, <strong>in</strong>troduc<strong>in</strong>g both Hamilton’s Pr<strong>in</strong>ciple of Stationary Action and Hamiltonian<br />

mechanics, marked the crown<strong>in</strong>g achievements for the development of variational pr<strong>in</strong>ciples <strong>in</strong> classical<br />

mechanics. A fundamental advantage of Hamiltonian mechanics is that it uses the conjugate coord<strong>in</strong>ates<br />

q p plus time , which is a considerable advantage <strong>in</strong> most branches of physics and eng<strong>in</strong>eer<strong>in</strong>g. Compared<br />

to Lagrangian mechanics, Hamiltonian mechanics has a significantly broader arsenal of powerful techniques<br />

that can be exploited to obta<strong>in</strong> an analytical solution of the <strong>in</strong>tegrals of the motion for complicated systems,<br />

as described <strong>in</strong> chapter 15. In addition, Hamiltonian dynamics provides a means of determ<strong>in</strong><strong>in</strong>g the<br />

unknown variables for which the solution assumes a soluble form, and is ideal for study of the fundamental<br />

underly<strong>in</strong>g physics <strong>in</strong> applications to fields such as quantum or statistical physics. As a consequence,<br />

Hamiltonian mechanics has become the preem<strong>in</strong>ent variational approach used <strong>in</strong> modern physics.<br />

This chapter has <strong>in</strong>troduced and discussed Hamilton’s Pr<strong>in</strong>ciple of Stationary Action, which underlies<br />

the elegant and remarkably powerful Lagrangian and Hamiltonian representations of algebraic mechanics.<br />

The basic concepts employed <strong>in</strong> algebraic mechanics are summarized below.<br />

(q ˙q) (91)<br />

(q ˙q) =0 (92)<br />

Hamilton’s Action Pr<strong>in</strong>ciple: As discussed <strong>in</strong> chapter 92, Hamiltonian mechanics is built upon Hamilton’s<br />

action functional<br />

Z <br />

(q p) =<br />

<br />

Hamilton’s Pr<strong>in</strong>ciple of least action states that<br />

Z <br />

(q p) =<br />

<br />

Generalized momentum :<br />

of the Lagrangian to be<br />

In chapter 72, the generalized (canonical) momentum was def<strong>in</strong>ed<strong>in</strong>terms<br />

≡<br />

(q ˙q)<br />

˙ <br />

Chapter 922 def<strong>in</strong>ed the generalized momentum <strong>in</strong> terms of the action functional to be<br />

(73)<br />

=<br />

(q p)<br />

<br />

(912)<br />

Generalized energy (q ˙ ): Jacobi’s Generalized Energy (q ˙ ) was def<strong>in</strong>ed<strong>in</strong>equation737 as<br />

(q ˙q) ≡ X µ<br />

<br />

(q ˙q)<br />

˙ − (q ˙q)<br />

(737)<br />

˙<br />

<br />

<br />

Hamiltonian function: (q p) The Hamiltonian (q p) was def<strong>in</strong>ed <strong>in</strong> terms of the generalized<br />

energy (q ˙q) plus the generalized momentum. That is<br />

(q p) ≡ (q ˙q)= X <br />

˙ − (q ˙q)=p · ˙q−(q ˙q)<br />

(737)<br />

where p q correspond to -dimensional vectors, e.g. q ≡ ( 1 2 ) and the scalar product p· ˙q = P ˙ .<br />

Chapter 82 used a Legendre transformation to derive this relation between the Hamiltonian and Lagrangian<br />

functions. Note that whereas the Lagrangian (q ˙q) is expressed <strong>in</strong> terms of the coord<strong>in</strong>ates q plus<br />

conjugate velocities ˙q, the Hamiltonian (q p) is expressed <strong>in</strong> terms of the coord<strong>in</strong>ates q plus their<br />

conjugate momenta p. For scleronomic systems, us<strong>in</strong>g the standard Lagrangian, <strong>in</strong> equations 744 and 729<br />

shows that the Hamiltonian simplifies to be equal to the total mechanical energy, that is, = + .


9.5. SUMMARY 233<br />

Generalized energy theorem: The equations of motion lead to the generalized energy theorem which<br />

statesthatthetimedependenceofthe Hamiltonian is related to the time dependence of the Lagrangian.<br />

(q p)<br />

= X "<br />

#<br />

X<br />

˙ (q ˙q)<br />

+ (q) − (738)<br />

<br />

<br />

<br />

<br />

Note that if all the generalized non-potential forces and Lagrange multiplier terms are zero, and if the<br />

Lagrangian is not an explicit function of time, then the Hamiltonian is a constant of motion.<br />

Lagrange equations of motion: Equation 660 gives that the Lagrange equations of motion are<br />

½ µ <br />

− ¾ X <br />

= (q)+ <br />

<br />

(660)<br />

˙ <br />

where =1 2 3 <br />

=1<br />

=1<br />

Hamilton’s equations of motion: Chapter 83 showed that a Legendre transform, plus the Lagrange-<br />

Euler equations, (964 965) lead to Hamilton’s equations of motion. Hamilton derived these equations of<br />

motion directly from the action functional, as shown <strong>in</strong> chapter 92<br />

(q p)<br />

<br />

˙ =<br />

(q p)<br />

<br />

˙ = − <br />

<br />

(q p)+<br />

(q ˙q)<br />

= −<br />

<br />

" <br />

X<br />

=1<br />

<br />

+ <br />

<br />

<br />

#<br />

(825)<br />

(826)<br />

(824)<br />

Note the symmetry of Hamilton’s two canonical equations. The canonical variables are treated<br />

as <strong>in</strong>dependent canonical variables Lagrange was the first to derive the canonical equations but he did not<br />

recognize them as a basic set of equations of motion. Hamilton derived the canonical equations of motion<br />

from his fundamental variational pr<strong>in</strong>ciple and made them the basis for a far-reach<strong>in</strong>g theory of dynamics.<br />

Hamilton’s equations give 2 first-order differential equations for for each of the degrees of freedom.<br />

Lagrange’s equations give second-order differential equations for the variables ˙ <br />

Hamilton-Jacobi equation:<br />

Jacobi equation.<br />

Hamilton used Hamilton’s Pr<strong>in</strong>ciple plus equation 919 to derive the Hamilton-<br />

<br />

<br />

+ (q p) =0<br />

(919)<br />

The solution of Hamilton’s equations is trivial if the Hamiltonian is a constant of motion, or when a set of<br />

generalized coord<strong>in</strong>ate can be identified for which all the coord<strong>in</strong>ates are constant, or are cyclic (also called<br />

ignorable coord<strong>in</strong>ates). Jacobi developed the mathematical framework of canonical transformation required<br />

to exploit the Hamilton-Jacobi equation.<br />

Hamilton’s Pr<strong>in</strong>ciple applied us<strong>in</strong>g <strong>in</strong>itial boundary conditions: The def<strong>in</strong>ition of Hamilton’s Pr<strong>in</strong>ciple<br />

assumes <strong>in</strong>tegration between the <strong>in</strong>itial time and f<strong>in</strong>al time . A recent development has extended<br />

applications of Hamilton’s Pr<strong>in</strong>ciple to apply to systems that are def<strong>in</strong>ed <strong>in</strong> terms of only the <strong>in</strong>itial boundary<br />

conditions. This method doubles the number of degrees of freedom and uses a coupl<strong>in</strong>g Lagrangian<br />

(q 2 ˙q 2 q 1 ˙q 1 ) between the correspond<strong>in</strong>g q 1 and q 2 doubled degrees of freedom<br />

<br />

˙ −<br />

− ∙ <br />

−<br />

=<br />

− − ¸<br />

<br />

˙ −<br />

≡ (q 1 ˙q 1 )<br />

(950)<br />

<br />

and where is a generalized nonconservative force derived from .


234 CHAPTER 9. HAMILTON’S ACTION PRINCIPLE<br />

Standard Lagrangians: Derivation of Lagrangian mechanics, us<strong>in</strong>g d’Alembert’s pr<strong>in</strong>ciple of virtual<br />

work, assumed that the Lagrangian is def<strong>in</strong>ed by equation 952<br />

(q ˙q) = ( ˙q) − (q)<br />

(952)<br />

Thiswasused<strong>in</strong>equation93 to derive the action <strong>in</strong> terms of the fundamental Lagrangian def<strong>in</strong>ed by equation<br />

952 The assumption that the action is the fundamental property <strong>in</strong>verts this procedure and now equation<br />

93 is used to derived the Lagrangian. That is, the assumption that Hamilton’s Pr<strong>in</strong>ciple is the foundation<br />

of algebraic mechanics def<strong>in</strong>es the Lagrangian <strong>in</strong> terms of the fundamental action <br />

Non-standard Lagrangians: The flexibility and power of Lagrangian mechanics can be extended to a<br />

broader range of dynamical systems by employ<strong>in</strong>g an extended def<strong>in</strong>ition of the Lagrangian that assumes that<br />

the action is the fundamental property, and then the Lagrangian is def<strong>in</strong>ed <strong>in</strong> terms of Hamilton’s variational<br />

action pr<strong>in</strong>ciple us<strong>in</strong>g equation 92. It was illustrated that the <strong>in</strong>verse variational calculus formalism can<br />

be used to identify non-standard Lagrangians that generate the required equations of motion. These nonstandard<br />

Lagrangians can be very different from the standard Lagrangian and do not separate <strong>in</strong>to k<strong>in</strong>etic<br />

and potential energy components. These alternative Lagrangians can be used to handle dissipative systems<br />

which are beyond the range of validity when us<strong>in</strong>g standard Lagrangians. That is, it was shown that several<br />

very different Lagrangians and Hamiltonians can be equivalent for generat<strong>in</strong>g useful equations of motion<br />

of a system. Currently the use of non-standard Lagrangians is a narrow, but active, frontier of classical<br />

mechanics with important applications to relativistic mechanics.<br />

Gauge <strong>in</strong>variance of the standard Lagrangian: It was shown that there is a cont<strong>in</strong>uum of equivalent<br />

standard Lagrangians that lead to the same set of equations of motion for a system. This feature is related<br />

to gauge <strong>in</strong>variance <strong>in</strong> mechanics. The follow<strong>in</strong>g transformations change the standard Lagrangian, but leave<br />

the equations of motion unchanged.<br />

1. The Lagrangian is <strong>in</strong>def<strong>in</strong>ite with respect to addition of a constant to the scalar potential which cancels<br />

out when the derivatives <strong>in</strong> the Euler-Lagrange differential equations are applied.<br />

2. Similarly the Lagrangian is <strong>in</strong>def<strong>in</strong>ite with respect to addition of a constant k<strong>in</strong>etic energy.<br />

3. The Lagrangian is <strong>in</strong>def<strong>in</strong>ite with respect to addition of a total time derivative of the form →<br />

+ [Λ( )] for any differentiable function Λ( ) of the generalized coord<strong>in</strong>ates, plus time, that has<br />

cont<strong>in</strong>uous second derivatives.<br />

Application of Hamilton’s Action Pr<strong>in</strong>ciple to mechanics: The derivation of the equations of motion<br />

for any system can be separated <strong>in</strong>to a hierarchical set of three stages <strong>in</strong> both sophistication and<br />

understand<strong>in</strong>g. <strong>Variational</strong> pr<strong>in</strong>ciples are employed dur<strong>in</strong>g the primary “action” stage and secondary “Hamilton/Lagrangian”<br />

stage to derive the required equations of motion, which then are solved dur<strong>in</strong>g the third<br />

“equations-of-motion stage”. Hamilton’s Action Pr<strong>in</strong>ciple, is a scalar function that is the basis for deriv<strong>in</strong>g<br />

the Lagrangian and Hamiltonian functions. The primary “action stage” uses Hamilton’s Action functional,<br />

= R <br />

<br />

(q ˙q) to derive the Lagrangian and Hamiltonian functionals that are based on Hamilton’s<br />

action functional and provide the most fundamental and sophisticated level of understand<strong>in</strong>g. The second<br />

“Hamiltonian/Lagrangian stage” <strong>in</strong>volves us<strong>in</strong>g the Lagrangian and Hamiltonian functionals to derive the<br />

equations of motion. The third “equations-of-motion stage” uses the derived equations of motion to solve<br />

for the motion subject to a given set of <strong>in</strong>itial boundary conditions. The Newtonian mechanics approach<br />

bypasses the primary “action” stage, as well as the secondary “Hamiltonian/Lagrangian” stage. That is,<br />

Newtonian mechanics starts at the third “equations-of-motion” stage, which does not allow exploit<strong>in</strong>g the<br />

considerable advantages provided by use of action, the Lagrangian, and the Hamiltonian. Newtonian mechanics<br />

requires that all the active forces be <strong>in</strong>cluded when deriv<strong>in</strong>g the equations of motion, which <strong>in</strong>volves<br />

deal<strong>in</strong>g with vector quantities. This is <strong>in</strong> contrast to the action, Lagrangian, and Hamiltonian which are<br />

scalar functionals. Both the primary “action” stage, and the secondary “Lagrangian/Hamiltonian” stage,<br />

exploit the powerful arsenal of mathematical techniques that have been developed for exploit<strong>in</strong>g variational<br />

pr<strong>in</strong>ciples.


Chapter 10<br />

Nonconservative systems<br />

10.1 Introduction<br />

Hamilton’s action pr<strong>in</strong>ciple, Lagrangian mechanics, and Hamiltonian mechanics, all exploit the concept of<br />

action which is a s<strong>in</strong>gle, <strong>in</strong>variant, quantity. These algebraic formulations of mechanics all are based on<br />

energy, which is a scalar quantity, and thus these formulations are easier to handle than the vector concept<br />

of force employed <strong>in</strong> Newtonian mechanics. Algebraic formulations provide a powerful and elegant approach<br />

to understand and develop the equations of motion of systems <strong>in</strong> nature. Chapters 6 − 9 applied variational<br />

pr<strong>in</strong>ciples to Hamilton’s action pr<strong>in</strong>ciple which led to the Lagrangian, and Hamiltonian formulations that<br />

simplify determ<strong>in</strong>ation of the equations of motion for systems <strong>in</strong> classical mechanics.<br />

A conservative force has the property that the total work done mov<strong>in</strong>g between two po<strong>in</strong>ts is <strong>in</strong>dependent<br />

of the taken path. That is, a conservative force is time symmetric and can be expressed <strong>in</strong> terms of the<br />

gradient of a scalar potential . Hamilton’s action pr<strong>in</strong>ciple implicitly assumes that the system is conservative<br />

for those degrees of freedom that are built <strong>in</strong>to the def<strong>in</strong>ition of the action, and the related Lagrangian, and<br />

Hamiltonian. The focus of this chapter is to discuss the orig<strong>in</strong>s of nonconservative motion and how it can<br />

be handled <strong>in</strong> algebraic mechanics.<br />

10.2 Orig<strong>in</strong>s of nonconservative motion<br />

Nonconservative degrees of freedom <strong>in</strong>volve irreversible processes, such as dissipation, damp<strong>in</strong>g, and also<br />

can result from course-gra<strong>in</strong><strong>in</strong>g, or ignor<strong>in</strong>g coupl<strong>in</strong>g to active degrees of freedom. The nonconservative role<br />

of ignored active degrees of freedom is illustrated by the weakly-coupled double harmonic oscillator system<br />

discussed below. Let the two harmonic oscillators have masses ( 1 2 ) uncoupled angular frequencies<br />

( 1 2 ) , and oscillation amplitudes ( 1 2 ). Assume that the coupl<strong>in</strong>g potential energy is = 1 2 The<br />

Lagrangian for this weakly-coupled double oscillator is<br />

( 1 2 ˙ 1 ˙ 2 )= 1<br />

2<br />

¡<br />

˙<br />

2<br />

1 − 2 1 2 1¢<br />

+ 1 2 + 2<br />

2<br />

¡<br />

˙<br />

2<br />

2 − 2 2 2 2¢<br />

(10.1)<br />

Note that the total Lagrangian is conservative s<strong>in</strong>ce the Lagrangian is explicitly time <strong>in</strong>dependent. As shown<br />

<strong>in</strong> chapter 142 the solution for the amplitudes of the oscillation for the coupled system are given by<br />

∙µ ¸ ∙µ ¸<br />

1 + 2 1 − 2<br />

1 () = s<strong>in</strong><br />

s<strong>in</strong><br />

<br />

(10.2)<br />

2<br />

2<br />

∙µ ¸ ∙µ ¸<br />

1 + 2 1 − 2<br />

2 () = cos<br />

cos<br />

<br />

(10.3)<br />

2<br />

2<br />

The system exhibits the common “beats” behavior where the coupled harmonic oscillators have an angular<br />

frequency that is the average oscillator frequency = ¡ 1+ 2<br />

¢<br />

2 and the oscillation <strong>in</strong>tensities are<br />

modulated at the difference frequency, = ¡ 1 − 2<br />

¢<br />

2 Although the total energy is conserved<br />

for this conservative system, this shared energy flows back and forth between the two coupled harmonic<br />

oscillators at the difference frequency. If the equations of motion for oscillator 1 ignore the coupl<strong>in</strong>g to the<br />

235


236 CHAPTER 10. NONCONSERVATIVE SYSTEMS<br />

motion of oscillator 2, that is, assume a constant average value 2 = h 2 i is used, then the <strong>in</strong>tensity | 1 | 2 and<br />

energy of the first oscillator still is modulated by the ¯¯s<strong>in</strong> ¡ 1 − 2<br />

¢<br />

2 ¯¯2 term. Thus the total energy for this<br />

truncated coupled-oscillator system is no longer conserved due to neglect of the energy flow<strong>in</strong>g<strong>in</strong>toandout<br />

of oscillator 1 due to its coupl<strong>in</strong>g to oscillator 2. That is, the solution for the truncated system of oscillator<br />

1 is not conservative s<strong>in</strong>ce it is exchang<strong>in</strong>g energy with the coupled, but ignored, second oscillator. This<br />

elementary example illustrates that ignor<strong>in</strong>g active degrees of freedom can transform a conservative system<br />

<strong>in</strong>to a nonconservative system, for which the equations of motion derived us<strong>in</strong>g the truncated Lagrangian is<br />

<strong>in</strong>correct.<br />

The above example illustrates the importance of <strong>in</strong>clud<strong>in</strong>g all active degrees of freedom when deriv<strong>in</strong>g the<br />

equations of motion, <strong>in</strong> order to ensure that the total system is conservative. Unfortunately, nonconservative<br />

systems due to viscous or frictional dissipation typically result from weak thermal <strong>in</strong>teractions with an<br />

enormous number of nearby atoms, which makes <strong>in</strong>clusion of all of these degrees of freedom impractical.<br />

Even though the detailed behavior of such dissipative degrees of freedom may not be of direct <strong>in</strong>terest, all<br />

the active degrees of freedom must be <strong>in</strong>cluded when apply<strong>in</strong>g Lagrangian or Hamiltonian mechanics.<br />

10.3 Algebraic mechanics for nonconservative systems<br />

S<strong>in</strong>ce Lagrangian and Hamiltonian formulations are <strong>in</strong>valid for the nonconservative degrees of freedom, the<br />

follow<strong>in</strong>g three approaches are used to <strong>in</strong>clude nonconservative degrees of freedom directly <strong>in</strong> the Lagrangian<br />

and Hamiltonian formulations of mechanics.<br />

1. Expand the number of degrees of freedom used to <strong>in</strong>clude all active degrees of freedom for the system,<br />

so that the expanded system is conservative. This is the preferred approach when it is viable. Hamilton’s<br />

action pr<strong>in</strong>ciple based on <strong>in</strong>itial conditions, <strong>in</strong>troduced <strong>in</strong> chapter 924, doubles the number of<br />

degrees of freedom, which can be used to account for the dissipative forces provid<strong>in</strong>g one approach to<br />

solve nonconservative systems. However, this approach typically is impractical for handl<strong>in</strong>g dissipated<br />

processes because of the large number of degrees of freedom that are <strong>in</strong>volved <strong>in</strong> thermal dissipation.<br />

2. Nonconservative forces can be <strong>in</strong>troduced directly at the equations of motion stage as generalized forces<br />

<br />

. This approach is used extensively. For the case of l<strong>in</strong>ear velocity dependence, the Rayleigh’s<br />

dissipation function provides an elegant and powerful way to express the generalized forces <strong>in</strong> terms of<br />

scalar potential energies.<br />

3. New degrees of freedom or effective forces can be postulated that are then <strong>in</strong>corporated <strong>in</strong>to the<br />

Lagrangian or the Hamiltonian <strong>in</strong> order to mimic the effects of the nonconservative forces.<br />

Examples that exploit the above three ways to <strong>in</strong>troduce nonconservative dissipative forces <strong>in</strong> algebraic<br />

formulations are given below.<br />

10.4 Rayleigh’s dissipation function<br />

As mentioned above, nonconservative systems <strong>in</strong>volv<strong>in</strong>g viscous or frictional dissipation, typically result from<br />

weak thermal <strong>in</strong>teractions with many nearby atoms, mak<strong>in</strong>g it impractical to <strong>in</strong>clude a complete set of active<br />

degrees of freedom. In addition, dissipative systems usually <strong>in</strong>volve complicated dependences on the velocity<br />

and surface properties that are best handled by <strong>in</strong>clud<strong>in</strong>g the dissipative drag force explicitly as a generalized<br />

drag force <strong>in</strong> the Euler-Lagrange equations. The drag force can have any functional dependence on velocity,<br />

position, or time.<br />

F = −( ˙q q)ˆv (10.4)<br />

Note that s<strong>in</strong>ce the drag force is dissipative the dom<strong>in</strong>ant component of the drag force must po<strong>in</strong>t <strong>in</strong> the<br />

opposite direction to the velocity vector.<br />

In 1881 Lord Rayleigh[Ray1881, Ray1887] showed that if a dissipative force F depends l<strong>in</strong>early on velocity,<br />

it can be expressed <strong>in</strong> terms of a scalar potential functional of the generalized coord<strong>in</strong>ates called the Rayleigh<br />

dissipation function R( ˙q). The Rayleigh dissipation function is an elegant way to <strong>in</strong>clude l<strong>in</strong>ear velocitydependent<br />

dissipative forces <strong>in</strong> both Lagrangian and Hamiltonian mechanics, as is illustrated below for both<br />

Lagrangian and Hamiltonian mechanics.


10.4. RAYLEIGH’S DISSIPATION FUNCTION 237<br />

10.4.1 Generalized dissipative forces for l<strong>in</strong>ear velocity dependence<br />

Consider equations of motion for the degrees of freedom, and assume that the dissipation depends l<strong>in</strong>early<br />

on velocity. Then, allow<strong>in</strong>g all possible cross coupl<strong>in</strong>g of the equations of motion for the equations of<br />

motion can be written <strong>in</strong> the form<br />

X<br />

[ ¨ + ˙ + − ()] = 0 (10.5)<br />

=1<br />

Multiply<strong>in</strong>g equation 105 by ˙ , take the time <strong>in</strong>tegral, and sum over , gives the follow<strong>in</strong>g energy equation<br />

X<br />

X<br />

Z <br />

¨ ˙ +<br />

X<br />

=1 =1 0<br />

=1 =1 0<br />

=1 =1 0<br />

X<br />

Z <br />

˙ ˙ +<br />

X<br />

X<br />

Z <br />

˙ =<br />

X<br />

Z <br />

0<br />

() ˙ (10.6)<br />

The right-hand term is the total energy supplied to the system by the external generalized forces ()<br />

at the time . The first time-<strong>in</strong>tegral term on the left-hand side is the total k<strong>in</strong>etic energy, while the third<br />

time-<strong>in</strong>tegral term equals the potential energy. The second <strong>in</strong>tegral term on the left is def<strong>in</strong>ed to equal 2R( ˙q)<br />

where Rayeigh’s dissipation function R( ˙q) is def<strong>in</strong>ed as<br />

R( ˙q)≡ 1 2<br />

X<br />

=1 =1<br />

X<br />

˙ ˙ (10.7)<br />

and the summations are over all particles of the system. This def<strong>in</strong>ition allows for complicated crosscoupl<strong>in</strong>g<br />

effects between the particles.<br />

The particle-particle coupl<strong>in</strong>g effects usually can be neglected allow<strong>in</strong>g use of the simpler def<strong>in</strong>ition that<br />

<strong>in</strong>cludes only the diagonal terms. Then the diagonal form of the Rayleigh dissipation function simplifies to<br />

R( ˙q)≡ 1 2<br />

X<br />

˙ 2 (10.8)<br />

Therefore the frictional force <strong>in</strong> the direction depends l<strong>in</strong>early on velocity ˙ ,thatis<br />

R( ˙q)<br />

<br />

= − = − ˙ (10.9)<br />

˙ <br />

In general, the dissipative force is the velocity gradient of the Rayleigh dissipation function,<br />

=1<br />

F = −∇ ˙q R( ˙q) (10.10)<br />

The physical significance of the Rayleigh dissipation function is illustrated by calculat<strong>in</strong>g the work done<br />

by one particle aga<strong>in</strong>st friction, which is<br />

Therefore<br />

= −F · r = −F · ˙q = ˙ 2 (10.11)<br />

2R( ˙q)= <br />

(10.12)<br />

<br />

which is the rate of energy (power) loss due to the dissipative forces <strong>in</strong>volved. The same relation is obta<strong>in</strong>ed<br />

after summ<strong>in</strong>g over all the particles <strong>in</strong>volved.<br />

Transform<strong>in</strong>g the frictional force <strong>in</strong>to generalized coord<strong>in</strong>ates requires equation 627<br />

ṙ = X <br />

Note that the derivative with respect to ˙ equals<br />

r <br />

˙ + r <br />

<br />

(10.13)<br />

ṙ <br />

˙ <br />

= r <br />

<br />

(10.14)


238 CHAPTER 10. NONCONSERVATIVE SYSTEMS<br />

Us<strong>in</strong>g equations 628 and 629 the component of the generalized frictional force <br />

is given by<br />

= X<br />

=1<br />

F · r <br />

<br />

=<br />

X<br />

=1<br />

F · ṙ <br />

˙ <br />

= −<br />

X<br />

=1<br />

∇ R( ˙q) · ṙ R( ˙q)<br />

= − (10.15)<br />

˙ ˙ <br />

Equation 1015 provides an elegant expression for the generalized dissipative force <br />

Rayleigh’s scalar dissipation potential R.<br />

10.4.2 Generalized dissipative forces for nonl<strong>in</strong>ear velocity dependence<br />

<strong>in</strong> terms of the<br />

The above discussion of the Rayleigh dissipation function was restricted to the special case of l<strong>in</strong>ear velocitydependent<br />

dissipation. Virga[Vir15] proposed that the scope of the classical Rayleigh-Lagrange formalism<br />

can be extended to <strong>in</strong>clude nonl<strong>in</strong>ear velocity dependent dissipation by assum<strong>in</strong>g that the nonconservative<br />

dissipative forces are def<strong>in</strong>ed by<br />

F ˙q)<br />

= −(q (10.16)<br />

˙q<br />

where the generalized Rayleigh dissipation function R(q ˙q) satisfies the general Lagrange mechanics relation<br />

<br />

− =0 (10.17)<br />

˙<br />

This generalized Rayleigh’s dissipation function elim<strong>in</strong>ates the prior restriction to l<strong>in</strong>ear dissipation processes,<br />

which greatly expands the range of validity for us<strong>in</strong>g Rayleigh’s dissipation function.<br />

10.4.3 Lagrange equations of motion<br />

L<strong>in</strong>ear dissipative forces can be directly, and elegantly, <strong>in</strong>cluded <strong>in</strong> Lagrangian mechanics by us<strong>in</strong>g Rayleigh’s<br />

dissipation function as a generalized force . Insert<strong>in</strong>g Rayleigh dissipation function 1015 <strong>in</strong> the generalized<br />

Lagrange equations of motion 660 gives<br />

½ µ <br />

− ¾ " <br />

#<br />

X <br />

= (q)+ R(q ˙q)<br />

− (10.18)<br />

˙ ˙ <br />

=1<br />

Where <br />

corresponds to the generalized forces rema<strong>in</strong><strong>in</strong>g after removal of the generalized l<strong>in</strong>ear, velocitydependent,<br />

frictional force . The holonomic forces of constra<strong>in</strong>t are absorbed <strong>in</strong>to the Lagrange multiplier<br />

term.<br />

10.4.4 Hamiltonian mechanics<br />

If the nonconservative forces depend l<strong>in</strong>early on velocity, and are derivable from Rayleigh’s dissipation<br />

function accord<strong>in</strong>g to equation 1015, then us<strong>in</strong>g the def<strong>in</strong>ition of generalized momentum gives<br />

"<br />

˙ = <br />

= <br />

<br />

#<br />

X <br />

+ (q)+ R(q ˙q)<br />

− (10.19)<br />

˙ ˙ <br />

=1<br />

" <br />

#<br />

(p q) X <br />

˙ = − + (q)+ R(q ˙q)<br />

− (10.20)<br />

˙ <br />

=1<br />

Thus Hamilton’s equations become<br />

˙ = <br />

(10.21)<br />

<br />

"<br />

˙ = − <br />

<br />

#<br />

X <br />

+ (q)+ R(q ˙q)<br />

− (10.22)<br />

˙ <br />

=1<br />

The Rayleigh dissipation function R(q ˙q) provides an elegant and convenient way to account for dissipative<br />

forces <strong>in</strong> both Lagrangian and Hamiltonian mechanics.


10.4. RAYLEIGH’S DISSIPATION FUNCTION 239<br />

10.1 Example: Driven, l<strong>in</strong>early-damped, coupled l<strong>in</strong>ear oscillators<br />

Consider the two identical, l<strong>in</strong>early damped, coupled<br />

oscillators (damp<strong>in</strong>g constant ) shown<strong>in</strong>thefigure. A<br />

periodic force = 0 cos() is applied to the left-hand<br />

mass . The k<strong>in</strong>etic energy of the system is<br />

The potential energy is<br />

= 1 2 ( ˙2 1 + ˙ 2 2)<br />

x 1<br />

x 2<br />

m<br />

Harmonically-driven, l<strong>in</strong>early-damped, coupled<br />

l<strong>in</strong>ear oscillators.<br />

m<br />

= 1 2 2 1 + 1 2 2 2 + 1 2 0 ( 2 − 1 ) 2 = 1 2 ( + 0 ) 2 1 + 1 2 ( + 0 ) 2 2 − 0 1 2<br />

Thus the Lagrangian equals<br />

= 1 ∙ 1<br />

2 ( ˙2 1 + ˙ 2 ) −<br />

2 ( + 0 ) 2 1 + 1 2 ( + 0 ) 2 2 − 0 1 2¸<br />

S<strong>in</strong>ce the damp<strong>in</strong>g is l<strong>in</strong>ear, it is possible to use the Rayleigh dissipation function<br />

The applied generalized forces are<br />

R = 1 2 ( ˙2 1 + ˙ 2 2)<br />

0 1 = cos () 0 2 =0<br />

Use the Euler-Lagrange equations 1018 to derive the equations of motion<br />

½ µ <br />

− ¾<br />

+ F<br />

X<br />

= 0 <br />

+ (q)<br />

˙ ˙ <br />

gives<br />

=1<br />

¨ 1 + ˙ 1 +( + 0 ) 1 − 0 2 = 0 cos ()<br />

¨ 2 + ˙ 2 +( + 0 ) 2 − 0 1 = 0<br />

These two coupled equations can be decoupled and simplified by mak<strong>in</strong>g a transformation to normal coord<strong>in</strong>ates,<br />

1 2 where<br />

1 = 1 − 2 2 = 1 + 2<br />

Thus<br />

1 = 1 2 ( 1 + 2 ) 2 = 1 2 ( 2 − 1 )<br />

Insert these <strong>in</strong>to the equations of motion gives<br />

(¨ 1 +¨ 2 )+( ˙ 1 + ˙ 2 )+( + 0 )( 1 + 2 ) − 0 ( 2 − 1 ) = 2 0 cos ()<br />

( 2 − 1 )+( 2 − 1 )+( + 0 )( 2 − 1 ) − 0 ( 1 + 2 ) = 0<br />

Add and subtract these two equations gives the follow<strong>in</strong>g two decoupled equations<br />

¨ 1 + ˙ 1 + ( +20 )<br />

1 = 0<br />

cos ()<br />

<br />

¨ 2 + ˙ 2 + 2 = 0<br />

cos ()<br />

<br />

q<br />

Def<strong>in</strong>e Γ = (+2<br />

1 =<br />

0 )<br />

<br />

2 = p <br />

= 0<br />

<br />

. Then the two <strong>in</strong>dependent equations of motion become<br />

¨ 1 + Γ˙ 1 + 2 1 1 = cos ()<br />

¨ 2 + Γ˙ 2 + 2 2 2 = cos ()


240 CHAPTER 10. NONCONSERVATIVE SYSTEMS<br />

This solution is a superposition of two <strong>in</strong>dependent, l<strong>in</strong>early-damped, driven normal modes 1 and 2 that<br />

have different natural frequencies 1 and 2 . For weak damp<strong>in</strong>g these two driven normal modesq<br />

each undergo<br />

damped oscillatory motion with the 1 and 2 normal modes exhibit<strong>in</strong>g resonances at 0 1 = 2 1 − 2 ¡ ¢<br />

Γ 2<br />

q<br />

2<br />

and 0 2 = 2 2 − 2 ¡ ¢<br />

Γ 2<br />

2<br />

10.2 Example: Kirchhoff’s rules for electrical circuits<br />

The mathematical equations govern<strong>in</strong>g the behavior of mechanical systems and electrical circuits<br />

have a close similarity. Thus variational methods can be used to derive the analogous behavior for electrical<br />

circuits. For example, for a system of separate circuits, the magnetic flux Φ through circuit due to<br />

electrical current = ˙ flow<strong>in</strong>g <strong>in</strong> circuit is given by<br />

Φ = ˙ <br />

where is the mutual <strong>in</strong>ductance. The diagonal term = corresponds to the self <strong>in</strong>ductance of<br />

circuit . The net magnetic flux Φ through circuit due to all circuits, is the sum<br />

X<br />

Φ = ˙ <br />

=1<br />

Thus the total magnetic energy which is analogous to k<strong>in</strong>etic energy is given by summ<strong>in</strong>g over all<br />

circuits to be<br />

= = 1 X X<br />

˙ ˙ <br />

2<br />

=1 =1<br />

Similarly the electrical energy stored <strong>in</strong> the mutual capacitance between the circuits, which<br />

is analogous to potential energy, is given by<br />

= = 1 X X <br />

2 <br />

=1 =1<br />

Thus the standard Lagrangian for this electric system is given by<br />

= − = 1 X X<br />

∙<br />

˙ ˙ − ¸<br />

<br />

2<br />

<br />

=1 =1<br />

Assum<strong>in</strong>g that Ohm’s Law is obeyed, that is, the dissipation force depends l<strong>in</strong>early on velocity, then the<br />

Rayleigh dissipation function can be written <strong>in</strong> the form<br />

R ≡ 1 X X<br />

˙ ˙ <br />

()<br />

2<br />

=1 =1<br />

where is the resistance matrix. Thus the dissipation force, expressed <strong>in</strong> volts, is given by<br />

= − R = 1 X<br />

˙ <br />

˙ 2<br />

Insert<strong>in</strong>g equations and <strong>in</strong>to equation 1018 plus mak<strong>in</strong>g the assumption that an additional generalized<br />

electrical force = () voltsisact<strong>in</strong>goncircuit then the Euler-Lagrange equations give the<br />

follow<strong>in</strong>g equations of motion.<br />

X<br />

∙<br />

¨ + ˙ + ¸<br />

<br />

= <br />

()<br />

<br />

=1<br />

This is a generalized version of Kirchhoff’s loop rule which can be seen by consider<strong>in</strong>g the case where the<br />

diagonal term = is the only non-zero term. Then<br />

∙<br />

¨ + ˙ + ¸<br />

<br />

= <br />

()<br />

<br />

This sum of the voltages is identical to the usual expression for Kirchhoff’s loop rule. This example<br />

illustrates the power of variational methods when applied to fields beyond classical mechanics.<br />

=1<br />

()<br />

()


10.5. DISSIPATIVE LAGRANGIANS 241<br />

10.5 Dissipative Lagrangians<br />

The prior discussion of nonconservative systems mentioned the follow<strong>in</strong>g three ways to <strong>in</strong>corporate dissipative<br />

processes <strong>in</strong>to Lagrangian or Hamiltonian mechanics. (1) Expand the number of degrees of freedom to <strong>in</strong>clude<br />

all the active dissipative active degrees of freedom as well as the conservative ones. (2) Use generalized forces<br />

to <strong>in</strong>corporate dissipative processes. (3) Add dissipative terms to the Lagrangian or Hamiltonian to mimic<br />

dissipation. The follow<strong>in</strong>g illustrates the use of dissipative Lagrangians.<br />

Bateman[Bat31] po<strong>in</strong>ted out that an isolated dissipative system is physically <strong>in</strong>complete, that is, a complete<br />

system must comprise at least two coupled subsystems where energy is transferred from a dissipat<strong>in</strong>g<br />

subsystem to an absorb<strong>in</strong>g subsystem. A complete system should comprise both the dissipat<strong>in</strong>g and absorb<strong>in</strong>g<br />

systems to ensure that the total system Lagrangian and Hamiltonian are conserved, as is assumed<br />

<strong>in</strong> conventional Lagrangian and Hamiltonian mechanics. Both Bateman and Dekker[Dek75] have illustrated<br />

that the equations of motion for a l<strong>in</strong>early-damped, free, one-dimensional harmonic oscillator are derivable<br />

us<strong>in</strong>g the Hamilton variational pr<strong>in</strong>ciple via <strong>in</strong>troduction of a fictitious complementary subsystem that mimics<br />

dissipative processes. The follow<strong>in</strong>g example illustrate that deriv<strong>in</strong>g the equations of motion for the<br />

l<strong>in</strong>early-damped, l<strong>in</strong>ear oscillator may be handled by three alternative equivalent non-standard Lagrangians<br />

that assume either: (1) a multidimensional system, (2) explicit time dependent Lagrangians and Hamiltonians,<br />

or (3) complex non-standard Lagrangians.<br />

10.3 Example: The l<strong>in</strong>early-damped, l<strong>in</strong>ear oscillator:<br />

Three toy dynamical models have been used to describe the l<strong>in</strong>early-damped, l<strong>in</strong>ear oscillator employ<strong>in</strong>g<br />

very different non-standard Lagrangians to generate the required Hamiltonians, and to derive the correct<br />

equations of motion.<br />

1: Dual-component Lagrangian: <br />

Bateman proposed a dual system compris<strong>in</strong>g a mass subject to two coupled one-dimensional variables<br />

( ) where is the observed variable and is the mirror variable for the subsystem that absorbs the energy<br />

dissipated by the subsystem .<br />

Assume a non-standard Lagrangian of the form<br />

= ∙<br />

˙ ˙ − Γ ¸<br />

2 2 [ ˙ − ˙] − 2 0<br />

()<br />

where Γ = is the damp<strong>in</strong>g coefficient. M<strong>in</strong>imiz<strong>in</strong>g by variation of the auxiliary variable , that is, Λ =0,<br />

leads to the uncoupled equation of motion for <br />

£¨ + Γ ˙ + <br />

2<br />

2<br />

0 ¤ =0<br />

()<br />

Similarly m<strong>in</strong>imiz<strong>in</strong>g by variation of the primary variable that is Λ =0 leads to the uncoupled equation<br />

of motion for <br />

£¨ − Γ ˙ + <br />

2<br />

2<br />

0 ¤ =0<br />

()<br />

Note that equation of motion () which was obta<strong>in</strong>ed by variation of the auxiliary variable corresponds<br />

to that for the usual free, l<strong>in</strong>early-damped, one-dimensional harmonic oscillator for the variable which<br />

dissipates energy as is discussed <strong>in</strong> chapter 35. The equation of motion () is obta<strong>in</strong>ed by variation of the<br />

primary variable and corresponds to a free l<strong>in</strong>ear, one-dimensional, oscillator for the variable that is<br />

absorb<strong>in</strong>g the energy dissipated by the dissipat<strong>in</strong>g system.<br />

The generalized momenta,<br />

≡ <br />

˙ <br />

can be used to derive the correspond<strong>in</strong>g Hamiltonian<br />

( )=[ ˙ + ˙ − ] = <br />

2 − Γ 2 [ − ]+ 2<br />

à µ ! 2 Γ<br />

2 0 − <br />

2<br />

()


242 CHAPTER 10. NONCONSERVATIVE SYSTEMS<br />

Note that this Hamiltonian is time <strong>in</strong>dependent, and thus is conserved for this complete dual-variable system.<br />

Us<strong>in</strong>g Hamilton’s equations of motion gives the same two uncoupled equations of motion as obta<strong>in</strong>ed us<strong>in</strong>g<br />

the Lagrangian, i.e. () and ().<br />

2: Time-dependent Lagrangian: <br />

The complementary subsystem of the above dual-component Lagrangian, that is added to the primary<br />

dissipative subsystem, is the adjo<strong>in</strong>t to the equations for the primary subsystem of <strong>in</strong>terest. In some cases, a<br />

set of the solutions of the complementary equations can be expressed <strong>in</strong> terms of the solutions of the primary<br />

subsystem allow<strong>in</strong>g the equations of motion to be expressed solely <strong>in</strong> terms of the variables of the primary<br />

subsystem. Inspection of the solutions of the damped harmonic oscillator, presented <strong>in</strong> chapter 35, implies<br />

that and must be related by the function<br />

= Γ<br />

()<br />

Therefore Bateman proposed a time-dependent, non-standard Lagrangian of the form<br />

= 2 Γ £ ˙ 2 − 2 0 2¤<br />

()<br />

This Lagrangian corresponds to a harmonic oscillator for which the mass = 0 Γ is accret<strong>in</strong>g<br />

exponentially with time <strong>in</strong> order to mimic the exponential energy dissipation. Use of this Lagrangian <strong>in</strong> the<br />

Euler-Lagrange equations gives the solution<br />

Γ £¨ + Γ ˙ + 2 0 ¤ =0<br />

()<br />

If the factor outside of the bracket is non-zero, then the equation <strong>in</strong> the bracket must be zero. The expression<br />

<strong>in</strong> the bracket is the required equation of motion for the l<strong>in</strong>early-damped l<strong>in</strong>ear oscillator. This Lagrangian<br />

generates a generalized momentum of<br />

= Γ ˙<br />

and the Hamiltonian is<br />

= ˙ − 2 = 2 <br />

2 −Γ + 2 2 0 Γ 2<br />

()<br />

The Hamiltonian is time dependent as expected. This leads to Hamilton’s equations of motion<br />

˙ = <br />

<br />

− ˙ = <br />

<br />

= <br />

−Γ<br />

()<br />

= 2 0 Γ ()<br />

Take the total time derivative of equation and use equation to substitute for ˙ gives<br />

Γ £¨ + Γ ˙ + 2 0 ¤ =0<br />

()<br />

If the term Γ is non-zero, then the term <strong>in</strong> brackets is zero. The term <strong>in</strong> the bracket is the usual equation<br />

of motion for the l<strong>in</strong>early-damped harmonic oscillator.<br />

3: Complex Lagrangian: <br />

Dekker proposed use of complex dynamical variables for solv<strong>in</strong>g the l<strong>in</strong>early-damped harmonic oscillator.<br />

It exploits the fact that, <strong>in</strong> pr<strong>in</strong>ciple, each second order differential equation can be expressed <strong>in</strong> terms of<br />

asetoffirst-order differential equations. This feature is the essential difference between Lagrangian and<br />

Hamiltonian mechanics. Let be complex and assume it can be expressed <strong>in</strong> the form of a real variable as<br />

= ˙ −<br />

µ<br />

+ Γ 2<br />

<br />

<br />

()<br />

Substitut<strong>in</strong>g this complex variable <strong>in</strong>to the relation<br />

∙<br />

˙ + + Γ ¸<br />

=0<br />

2<br />

()<br />

leads to the second-order equation for the real variable of<br />

¨ + Γ ˙ + 2 0 =0<br />

()


10.6. SUMMARY 243<br />

This is the desired equation of motion for the l<strong>in</strong>early-damped harmonic oscillator. This result also can be<br />

shown by tak<strong>in</strong>g the time derivative of equation () and tak<strong>in</strong>g only the real part, i.e.<br />

¨ + ˙ + Γ µ<br />

2 ˙ =¨ + − Γ <br />

˙ + Γ ˙ =¨ + Γ ˙ + 2<br />

2<br />

0 =0<br />

()<br />

This feature is exploited us<strong>in</strong>g the follow<strong>in</strong>g Lagrangian<br />

= 2 (∗ ˙ − ˙ ∗ ) −<br />

∙<br />

− Γ ¸<br />

∗ <br />

2<br />

()<br />

where 2 ≡ 2 0 − ¡ ¢<br />

Γ 2.<br />

2 The Lagrangian is real for a conservative system and complex for a<br />

dissipative system. Us<strong>in</strong>g the Lagrange-Euler equation for variation of ∗ , that is, Λ ∗ =0, gives<br />

equation () which leads to the required equation of motion ()<br />

The canonical conjugate momenta are given by<br />

= <br />

˙<br />

˜ = <br />

˙ ∗<br />

The above Lagrangian plus canonically conjugate momenta lead to the complimentary Hamiltonians<br />

µ<br />

( ˜ ∗ ) = + Γ <br />

(˜∗ ∗ − )<br />

2<br />

µ<br />

˜ ( ˜ ∗ ) = − Γ <br />

(˜∗ ∗ − )<br />

2<br />

()<br />

()<br />

()<br />

These Hamiltonians give Hamilton equations of motion that lead to the correct equations of motion for <br />

and ∗<br />

The above examples have shown that three very different, non-standard, Lagrangians, plus their correspond<strong>in</strong>g<br />

Hamiltonians, all lead to the correct equation of motion for the l<strong>in</strong>early-damped harmonic oscillator.<br />

This illustrates the power of us<strong>in</strong>g non-standard Lagrangians to describe dissipative motion <strong>in</strong> classical<br />

mechanics. However, postulat<strong>in</strong>g non-standard Lagrangians to produce the required equations of motion<br />

appears to be of questionable usefulness. A fundamental approach is needed to build a firm foundation upon<br />

which non-standard Lagrangian mechanics can be based. Non-standard Lagrangian mechanics rema<strong>in</strong>s an<br />

active, albeit narrow, frontier of classical mechanics<br />

10.6 Summary<br />

Dissipative drag forces are non-conservative and usually are velocity dependent. Chapter 4 showed that the<br />

motion of non-l<strong>in</strong>ear dissipative dynamical systems can be highly sensitive to the <strong>in</strong>itial conditions and can<br />

lead to chaotic motion.<br />

Algebraic mechanics for nonconservative systems S<strong>in</strong>ce Lagrangian and Hamiltonian formulations<br />

are <strong>in</strong>valid for the nonconservative degrees of freedom, the follow<strong>in</strong>g three approaches are used to <strong>in</strong>clude<br />

nonconservative degrees of freedom directly <strong>in</strong> the Lagrangian and Hamiltonian formulations of mechanics.<br />

1. Expand the number of degrees of freedom used to <strong>in</strong>clude all active degrees of freedom for the system,<br />

so that the expanded system is conservative. This is the preferred approach when it is viable. Unfortunately<br />

this approach typically is impractical for handl<strong>in</strong>g dissipated processes because of the large<br />

number of degrees of freedom that are <strong>in</strong>volved <strong>in</strong> thermal dissipation.<br />

2. Nonconservative forces can be <strong>in</strong>troduced directly at the equations of motion stage as generalized forces<br />

<br />

. This approach is used extensively. For the case of l<strong>in</strong>ear velocity dependence, the Rayleigh’s<br />

dissipation function provides an elegant and powerful way to express the generalized forces <strong>in</strong> terms of<br />

scalar potential energies.<br />

3. New degrees of freedom or effective forces can be postulated that are then <strong>in</strong>corporated <strong>in</strong>to the<br />

Lagrangian or the Hamiltonian <strong>in</strong> order to mimic the effects of the nonconservative forces.


244 CHAPTER 10. NONCONSERVATIVE SYSTEMS<br />

Rayleigh’s dissipation function Generalized dissipative forces that have a l<strong>in</strong>ear velocity dependence<br />

can be easily handled <strong>in</strong> Lagrangian or Hamiltonian mechanics by <strong>in</strong>troduc<strong>in</strong>g the powerful Rayleigh’s<br />

dissipation function R( ˙q) where<br />

R( ˙q)≡ 1 X X<br />

˙ ˙ <br />

(107)<br />

2<br />

=1 =1<br />

This approach is used extensively <strong>in</strong> physics. This approach has been generalized by def<strong>in</strong><strong>in</strong>g a l<strong>in</strong>ear velocity<br />

dependent Rayleigh dissipation function<br />

F ˙q)<br />

<br />

= −(q (1016)<br />

˙q<br />

where the generalized Rayleigh dissipation function R(q ˙q) satisfies the general Lagrange mechanics relation<br />

<br />

− <br />

˙ =0<br />

(1017)<br />

This generalized Rayleigh’s dissipation function elim<strong>in</strong>ates the prior restriction to l<strong>in</strong>ear dissipation processes,<br />

which greatly expands the range of validity for us<strong>in</strong>g Rayleigh’s dissipation function.<br />

Rayleigh dissipation <strong>in</strong> Lagrange equations of motion L<strong>in</strong>ear dissipative forces can be directly, and<br />

elegantly, <strong>in</strong>cluded <strong>in</strong> Lagrangian mechanics by us<strong>in</strong>g Rayleigh’s dissipation function as a generalized force<br />

. Insert<strong>in</strong>g Rayleigh dissipation function 1015 <strong>in</strong> the generalized Lagrange equations of motion 660 gives<br />

½ µ <br />

− ¾ " <br />

#<br />

X <br />

= (q)+ R(q ˙q)<br />

− (1018)<br />

˙ ˙ <br />

=1<br />

Where <br />

corresponds to the generalized forces rema<strong>in</strong><strong>in</strong>g after removal of the generalized l<strong>in</strong>ear, velocitydependent,<br />

frictional force <br />

. The holonomic forces of constra<strong>in</strong>t are absorbed <strong>in</strong>to the Lagrange multiplier<br />

term.<br />

Rayleigh dissipation <strong>in</strong> Hamiltonian mechanics If the nonconservative forces depend l<strong>in</strong>early on<br />

velocity, and are derivable from Rayleigh’s dissipation function accord<strong>in</strong>g to equation 1015, then us<strong>in</strong>g the<br />

def<strong>in</strong>ition of generalized momentum gives<br />

"<br />

˙ = <br />

= <br />

<br />

#<br />

X <br />

+ (q)+ R(q ˙q)<br />

− (1019)<br />

˙ ˙ <br />

=1<br />

" <br />

#<br />

(p q) X <br />

˙ = − + (q)+ R(q ˙q)<br />

− (1020)<br />

˙ <br />

Thus Hamilton’s equations become<br />

=1<br />

˙ = <br />

<br />

˙ = − <br />

<br />

+<br />

"<br />

X <br />

#<br />

<br />

(q)+ <br />

−<br />

<br />

=1<br />

R(q ˙q)<br />

˙ <br />

(1021)<br />

(1022)<br />

The Rayleigh dissipation function R(q ˙q) provides an elegant and convenient way to account for dissipative<br />

forces <strong>in</strong> both Lagrangian and Hamiltonian mechanics.<br />

Dissipative Lagrangians or Hamiltonians New degrees of freedom or effective forces can be postulated<br />

that are then <strong>in</strong>corporated <strong>in</strong>to the Lagrangian or the Hamiltonian <strong>in</strong> order to mimic the effects of the<br />

nonconservative forces. This approach has been used for special cases.


Chapter 11<br />

Conservative two-body central forces<br />

11.1 Introduction<br />

Conservative two-body central forces are important <strong>in</strong> physics because of the pivotal role that the Coulomb<br />

and the gravitational forces play <strong>in</strong> nature. The Coulomb force plays a role <strong>in</strong> electrodynamics, molecular,<br />

atomic, and nuclear physics, while the gravitational force plays an analogous role <strong>in</strong> celestial mechanics.<br />

Therefore this chapter focusses on the physics of systems <strong>in</strong>volv<strong>in</strong>g conservative two-body central forces<br />

because of the importance and ubiquity of these conservative two-body central forces <strong>in</strong> nature.<br />

A conservative two-body central force has the follow<strong>in</strong>g three important attributes.<br />

1. Conservative: A conservative force depends only on the particle position, that is, the force is not<br />

time dependent. Moreover the work done by the force mov<strong>in</strong>g a body between any two po<strong>in</strong>ts 1 and 2<br />

is path <strong>in</strong>dependent. Conservative fields are discussed <strong>in</strong> chapter 210.<br />

2. Two-body: A two-body force between two bodies depends only on the relative locations of the two<br />

<strong>in</strong>teract<strong>in</strong>g bodies and is not <strong>in</strong>fluenced by the proximity of additional bodies. For two-body forces<br />

act<strong>in</strong>g between bodies, the force on body 1 is the vector superposition of the two-body forces due<br />

to the <strong>in</strong>teractions with each of the other − 1 bodies. This differs from three-body forces where the<br />

force between any two bodies is <strong>in</strong>fluenced by the proximity of a third body.<br />

3. Central: A central force field depends on the distance 12 from the orig<strong>in</strong> of the force at po<strong>in</strong>t 1 to<br />

the body location at po<strong>in</strong>t 2, and the force is directed along the l<strong>in</strong>e jo<strong>in</strong><strong>in</strong>g them, that is, ˆr 12 .<br />

A conservative, two-body, central force comb<strong>in</strong>es the above three attributes and can be expressed as,<br />

F 21 =( 12 )ˆr 12<br />

(11.1)<br />

The force field F 21 has a magnitude ( 12 ) that depends only on the magnitude of the relative separation<br />

vector r 12 = r 2 − r 1 between the orig<strong>in</strong> of the force at po<strong>in</strong>t 1 and po<strong>in</strong>t 2 wheretheforceacts,andtheforce<br />

is directed along the l<strong>in</strong>e jo<strong>in</strong><strong>in</strong>g them, that is, ˆr 12 .<br />

Chapter 210 showed that if a two-body central force is conservative, then it can be written as the gradient<br />

of a scalar potential energy () which is a function of the distance from the center of the force field.<br />

F 21 = −∇( 12 ) (11.2)<br />

As discussed <strong>in</strong> chapter 2, the ability to represent the conservative central force by a scalar function ()<br />

greatly simplifies the treatment of central forces.<br />

The Coulomb and gravitational forces both are true conservative, two-body, central forces whereas the<br />

nuclear force between nucleons <strong>in</strong> the nucleus has three-body components. Two bodies <strong>in</strong>teract<strong>in</strong>g via a<br />

two-body central force is the simplest possible system to consider, but equation 111 is applicable equally<br />

for bodies <strong>in</strong>teract<strong>in</strong>g via two-body central forces because the superposition pr<strong>in</strong>ciple applies for two-body<br />

central forces. This chapter will focus first on the motion of two bodies <strong>in</strong>teract<strong>in</strong>g via conservative two-body<br />

central forces followed by a brief discussion of the motion for 2 <strong>in</strong>teract<strong>in</strong>g bodies.<br />

245


246 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

11.2 Equivalent one-body representation for two-body motion<br />

The motion of two bodies, 1 and 2, <strong>in</strong>teract<strong>in</strong>gviatwo-body<br />

central forces, requires 6 spatial coord<strong>in</strong>ates, that is, three each<br />

for r 1 and r 2 . S<strong>in</strong>ce the two-body central force only depends on<br />

the relative separation r = r 1 − r 2 of the two bodies, it is more<br />

convenient to separate the 6 degrees of freedom <strong>in</strong>to 3 spatial<br />

coord<strong>in</strong>ates of relative motion r plus 3 spatial coord<strong>in</strong>ates for<br />

the center-of-mass location R asdescribed<strong>in</strong>chapter27. It will<br />

be shown here that the equation of motion for relative motion<br />

of the two-bodies <strong>in</strong> the center of mass can be represented by an<br />

equivalent one-body problem which simplifies the mathematics.<br />

Consider two bodies acted upon by a conservative two-body<br />

central force, where the position vectors r 1 and r 2 specify the<br />

location of each particle as illustrated <strong>in</strong> figure 111. Analternate<br />

set of six variables would be the three components of the center<br />

of mass position vector R and the three components specify<strong>in</strong>g<br />

the difference vector r def<strong>in</strong>ed by figure 111. Def<strong>in</strong>e the vectors<br />

r 0 1 and r 0 2 as the position vectors of the masses 1 and 2 with<br />

respect to the center of mass. Then<br />

r 1 = R + r 0 1 (11.3)<br />

r 2 = R + r 0 2<br />

Figure 11.1: Center of mass cord<strong>in</strong>ates for<br />

the two-body system.<br />

By the def<strong>in</strong>ition of the center of mass<br />

and<br />

R = 1r 1 + 2 r 2<br />

1 + 2<br />

(11.4)<br />

1 r 0 1 + 2 r 0 2 =0 (11.5)<br />

so that<br />

Therefore<br />

that is,<br />

Similarly;<br />

Substitut<strong>in</strong>g these <strong>in</strong>to equation 113 gives<br />

− 1<br />

2<br />

r 0 1 = r 0 2 (11.6)<br />

r = r 0 1 − r 0 2 = 1 + 2<br />

2<br />

r 0 1 (11.7)<br />

r 0 1 = 2<br />

1 + 2<br />

r (11.8)<br />

r 0 2 = − 1<br />

1 + 2<br />

r (11.9)<br />

r 1 = R + r 0 1 = R + 2<br />

r<br />

1 + 2<br />

r 2 = R + r 0 2 = R − 1<br />

r (11.10)<br />

1 + 2<br />

That is, the two vectors r 1 r 2 are written <strong>in</strong> terms of the position vector for the center of mass R and the<br />

position vector r for relative motion <strong>in</strong> the center of mass.<br />

Assum<strong>in</strong>g that the two-body central force is conservative and represented by (), then the Lagrangian<br />

of the two-body system can be written as<br />

= 1 2 1 |ṙ 1 | 2 + 1 2 2 |ṙ 2 | 2 − () (11.11)


11.2. EQUIVALENT ONE-BODY REPRESENTATION FOR TWO-BODY MOTION 247<br />

Differentiat<strong>in</strong>g equations 1110 with respect to time, and <strong>in</strong>sert<strong>in</strong>g them <strong>in</strong>to the Lagrangian, gives<br />

= 1 ¯¯¯Ṙ<br />

2 ¯<br />

¯2 + 1 2 |ṙ|2 − () (11.12)<br />

where the total mass is def<strong>in</strong>ed as<br />

= 1 + 2 (11.13)<br />

and the reduced mass is def<strong>in</strong>ed by<br />

≡ 1 2<br />

(11.14)<br />

1 + 2<br />

or equivalently<br />

1<br />

= 1 + 1 (11.15)<br />

1 2<br />

The total Lagrangian can be separated <strong>in</strong>to two <strong>in</strong>dependent parts<br />

= 1 ¯¯¯Ṙ<br />

2 ¯<br />

¯2 + (11.16)<br />

where<br />

= 1 2 |ṙ|2 − () (11.17)<br />

Assum<strong>in</strong>g that no external forces are act<strong>in</strong>g, then <br />

R<br />

=0and the three Lagrange equations for each of the<br />

three coord<strong>in</strong>ates of the R coord<strong>in</strong>ate can be written as<br />

<br />

Ṙ = P <br />

=0 (11.18)<br />

<br />

That is, for a pure central force, the center-of-mass momentum P is a constant of motion where<br />

P = = Ṙ (11.19)<br />

Ṙ<br />

It is convenient to work <strong>in</strong> the center-of-mass frame us<strong>in</strong>g<br />

the effective Lagrangian . In the center-of-mass frame of<br />

reference, the translational k<strong>in</strong>etic energy 1 2<br />

¯¯¯Ṙ ¯<br />

¯2<br />

associated<br />

with center-of-mass motion is ignored, and only the energy <strong>in</strong><br />

the center-of-mass is considered. This center-of-mass energy<br />

is the energy <strong>in</strong>volved <strong>in</strong> the <strong>in</strong>teraction between the collid<strong>in</strong>g<br />

bodies. Thus, <strong>in</strong> the center-of-mass, the problem has been reduced<br />

to an equivalent one-body problem of a mass mov<strong>in</strong>g<br />

about a fixedforcecenterwithapathgivenbyr which is the<br />

separation vector between the two bodies, as shown <strong>in</strong> figure<br />

112. In reality, both masses revolve around their center of<br />

mass, also called the barycenter, <strong>in</strong> the center-of-mass frame<br />

as shown <strong>in</strong> figure 112. Know<strong>in</strong>g r allows the trajectory of<br />

each mass about the center of mass r 0 1 and r 0 2 to be calculated.<br />

Of course the true path <strong>in</strong> the laboratory frame of<br />

reference must take <strong>in</strong>to account both the translational motion<br />

of the center of mass, <strong>in</strong> addition to the motion of the<br />

equivalent one-body representation relative to the barycenter.<br />

Be careful to remember the difference between the actual trajectories<br />

of each body, and the effective trajectory assumed<br />

when us<strong>in</strong>g the reduced mass which only determ<strong>in</strong>es the relative<br />

separation r of the two bodies. This reduction to an<br />

equivalent one-body problem greatly simplifies the solution<br />

Figure 11.2: Orbits of a two-body system<br />

with mass ratio of 2 rotat<strong>in</strong>g about the<br />

center-of-mass, O. The dashed ellipse is the<br />

equivalent one-body orbit with the center of<br />

force at the focus O.<br />

of the motion, but it misrepresents the actual trajectories and the spatial locations of each mass <strong>in</strong> space.<br />

The equivalent one-body representation will be used extensively throughout this chapter.


248 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

11.3 Angular momentum L<br />

The notation used for the angular momentum vector is L where the magnitude is designated by |L| = .<br />

Be careful not to confuse the angular momentum vector L with the Lagrangian Note that the angular<br />

momentum for two-body rotation about the center of mass with angular velocity is identical when evaluated<br />

<strong>in</strong> either the laboratory or equivalent two-body representation. That is, us<strong>in</strong>g equations 118 and 119<br />

L = m 1 1 02 ω + m 2 2 02 ω = 2 ω (11.20)<br />

The center-of-mass Lagrangian leads to the follow<strong>in</strong>g two general properties regard<strong>in</strong>g the angular momentum<br />

vector L.<br />

1) The motion lies entirely <strong>in</strong> a plane perpendicular to the fixed direction of the total angular momentum<br />

vector. This is because<br />

L · r = r × p · r =0 (11.21)<br />

that is, the radius vector is <strong>in</strong> the plane perpendicular to the total angular momentum vector. Thus, it is<br />

possible to express the Lagrangian <strong>in</strong> polar coord<strong>in</strong>ates, ( ) rather than spherical coord<strong>in</strong>ates. In polar<br />

coord<strong>in</strong>ates the center-of-mass Lagrangian becomes<br />

= 1 ³<br />

2 ˙ 2 + ˙2´ 2 − () (11.22)<br />

2) If the potential is spherically symmetric, then the polar angle is cyclic and therefore Noether’s<br />

theorem gives that the angular momentum p ≡ L = r × p is a constant of motion. That is, s<strong>in</strong>ce <br />

=0<br />

then the Lagrange equations imply that<br />

ṗ = <br />

=0 (11.23)<br />

˙ψ<br />

where the vectors ṗ and ˙ψ imply that equation 1123 refers to three <strong>in</strong>dependent equations correspond<strong>in</strong>g<br />

to the three components of these vectors. Thus the angular momentum p conjugate to ψ is a constant of<br />

motion. The generalized momentum p is a first <strong>in</strong>tegral of the motion which equals<br />

p = <br />

˙ψ = 2 ˙ψ = ˆp (11.24)<br />

where the magnitude of the angular momentum , and the direction ˆp both are constants of motion.<br />

A simple geometric <strong>in</strong>terpretation of equation 1124 is illustrated<br />

<strong>in</strong> figure 113 The radius vector sweeps out an area A<br />

<strong>in</strong> time where<br />

y<br />

A = 1 r × v (11.25)<br />

2<br />

and the vector A is perpendicular to the − plane. The rate<br />

of change of area is<br />

A<br />

= 1 2 r × v (11.26)<br />

But the angular momentum is<br />

r+dr<br />

r<br />

L = r × p = r × v =2 A (11.27)<br />

<br />

Thus the conservation of angular momentum implies that the<br />

areal velocity <br />

<br />

also is a constant of motion This fact is called<br />

Kepler’s second law of planetary motion which he deduced <strong>in</strong><br />

1609 basedonTychoBrahe’s55 years of observational records<br />

x<br />

O<br />

of the motion of Mars. Kepler’s second law implies that a<br />

planet moves fastest when closest to the sun and slowest when<br />

farthest from the sun. Note that Kepler’s second law is a statement<br />

of the conservation of angular momentum which is <strong>in</strong>dependent<br />

of the radial form of the central potential.<br />

Figure 11.3: Area swept out by the radius<br />

vector<strong>in</strong>thetimedt.


11.4. EQUATIONS OF MOTION 249<br />

11.4 Equations of motion<br />

The equations of motion for two bodies <strong>in</strong>teract<strong>in</strong>g via a conservative two-body central force can be determ<strong>in</strong>ed<br />

us<strong>in</strong>g the center of mass Lagrangian, given by equation 1122 For the radial coord<strong>in</strong>ate, the<br />

operator equation Λ =0for Lagrangian mechanics leads to<br />

But<br />

therefore the radial equation of motion is<br />

<br />

( ˙) − ˙ 2 + <br />

<br />

˙ =<br />

=0 (11.28)<br />

<br />

2 (11.29)<br />

¨ = − <br />

+ 2<br />

3 (11.30)<br />

Similarly, for the angular coord<strong>in</strong>ate, the operator equation Λ =0leads to equation 1124. That is,<br />

the angular equation of motion for the magnitude of is<br />

= <br />

˙ = 2 ˙ = (11.31)<br />

Lagrange’s equations have given two equations of motion, one dependent on radius and the other on<br />

the polar angle . Note that the radial acceleration is just a statement of Newton’s Laws of motion for the<br />

radial force <strong>in</strong> the center-of-mass system of<br />

This can be written <strong>in</strong> terms of an effective potential<br />

() ≡ ()+<br />

2<br />

2 2 (11.33)<br />

which leads to an equation of motion<br />

= − <br />

+ 2<br />

3 (11.32)<br />

= ¨ = − ()<br />

<br />

(11.34)<br />

S<strong>in</strong>ce<br />

2<br />

<br />

= ˙ 2 , the second term <strong>in</strong> equation (1133)<br />

3<br />

is the usual centrifugal force that orig<strong>in</strong>ates because the<br />

variable is <strong>in</strong> a non-<strong>in</strong>ertial, rotat<strong>in</strong>g frame of reference.<br />

Note that the angular equation of motion is <strong>in</strong>dependent<br />

of the radial dependence of the conservative two-body<br />

central force.<br />

Figure 114 shows, by dashed l<strong>in</strong>es, the radial dependence<br />

of the potential correspond<strong>in</strong>g to the attractive<br />

<strong>in</strong>verse square law force, that is = − ,andthepotential<br />

correspond<strong>in</strong>g to the centrifugal term<br />

2<br />

2<br />

correspond<strong>in</strong>g<br />

to a repulsive centrifugal force. The sum of<br />

2<br />

these two potentials (), shown by the solid l<strong>in</strong>e,<br />

has a m<strong>in</strong>imum m<strong>in</strong> value at a certa<strong>in</strong> radius similar<br />

to that manifest by the diatomic molecule discussed <strong>in</strong><br />

example 27.<br />

It is remarkable that the six-dimensional equations<br />

of motion, for two bodies <strong>in</strong>teract<strong>in</strong>g via a two-body<br />

Figure 11.4: The attractive <strong>in</strong>verse-square law potential<br />

( 2<br />

<br />

), the centrifugal potential (<br />

2<br />

),and<br />

2<br />

the comb<strong>in</strong>ed effective bound potential.<br />

central force, has been reduced to trivial center-of-mass translational motion, plus a one-dimensional onebody<br />

problem given by (1134) <strong>in</strong> terms of the relative separation and an effective potential ().


250 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

11.5 Differential orbit equation:<br />

The differential orbit equation relates the shape of the orbital motion, <strong>in</strong> plane polar coord<strong>in</strong>ates, to the<br />

radial dependence of the two-body central force. A B<strong>in</strong>et coord<strong>in</strong>ate transformation, which depends on the<br />

functional form of F(r) can simplify the differential orbit equation. For the <strong>in</strong>verse-square law force, the<br />

best B<strong>in</strong>et transformed variable is which is def<strong>in</strong>ed to be<br />

≡ 1 <br />

(11.35)<br />

Insert<strong>in</strong>g the transformed variable <strong>in</strong>to equation 1129 gives<br />

˙ = 2<br />

(11.36)<br />

<br />

From the def<strong>in</strong>ition of the new variable<br />

<br />

= −−2 = −−2<br />

˙ = − <br />

(11.37)<br />

<br />

Differentiat<strong>in</strong>g aga<strong>in</strong> gives<br />

2 <br />

2 = − µ µ 2<br />

2 <br />

= −<br />

2 (11.38)<br />

Substitut<strong>in</strong>g these <strong>in</strong>to Lagrange’s radial equation of motion gives<br />

2 <br />

2 + = − 1<br />

2 2 ( 1 ) (11.39)<br />

B<strong>in</strong>et’s differential orbit equation directly relates and which determ<strong>in</strong>es the overall shape of the orbit<br />

trajectory. This shape is crucial for understand<strong>in</strong>g the orbital motion of two bodies <strong>in</strong>teract<strong>in</strong>g via a twobody<br />

central force. Note that for the special case of an <strong>in</strong>verse square-law force, that is where ( 1 )=2 ,<br />

then the right-hand side of equation 1139 equals a constant − <br />

<br />

s<strong>in</strong>ce the orbital angular momentum is a<br />

2<br />

conserved quantity.<br />

11.1 Example: Central force lead<strong>in</strong>g to a circular orbit =2 cos <br />

B<strong>in</strong>et’s differential orbit equation can be used to derive the<br />

central potential that leads to the assumed circular trajectory<br />

of =2 cos where is the radius of the circular orbit.<br />

Note that this circular orbit passes through the orig<strong>in</strong> of the<br />

central force when =2 cos =0<br />

Insert<strong>in</strong>g this trajectory <strong>in</strong>to B<strong>in</strong>et’s differential orbit equation<br />

1139 gives<br />

1 2 (cos ) −1<br />

2 2 + 1<br />

2 (cos )−1 = − 2 42 (cos ) 2 ( 1 ) ()<br />

Note that the differential is given by<br />

2 (cos ) −1<br />

2 = µ s<strong>in</strong> <br />

cos 3 = 2s<strong>in</strong>2 <br />

cos 3 + 1<br />

cos <br />

Insert<strong>in</strong>g this differential <strong>in</strong>to equation gives<br />

2s<strong>in</strong> 2 <br />

cos 3 + 1<br />

cos + 1<br />

cos = 2<br />

cos 3 = − 2 83 (cos ) 2 ( 1 )<br />

Thus the radial dependence of the required central force is<br />

= −<br />

2 2<br />

8 3 cos 5 = 2<br />

−82 <br />

r<br />

Circular trajectory pass<strong>in</strong>g through the<br />

orig<strong>in</strong> of the central force.<br />

1<br />

5 = − 5<br />

This corresponds to an attractive central force that depends to the fifth power on the <strong>in</strong>verse radius r. Note<br />

that this example is unrealistic s<strong>in</strong>ce the assumed orbit implies that the potential and k<strong>in</strong>etic energies are<br />

<strong>in</strong>f<strong>in</strong>ite when → 0 at → 2 .<br />

R


11.6. HAMILTONIAN 251<br />

11.6 Hamiltonian<br />

S<strong>in</strong>ce the center-of-mass Lagrangian is not an explicit function of time, then<br />

<br />

<br />

= − <br />

<br />

=0 (11.40)<br />

Thus the center-of mass Hamiltonian is a constant of motion. However, s<strong>in</strong>ce the transformation to<br />

center of mass can be time dependent, then 6= that is, it does not <strong>in</strong>clude the total energy because<br />

the k<strong>in</strong>etic energy of the center-of-mass motion has been omitted from . Also, s<strong>in</strong>ce no transformation<br />

is <strong>in</strong>volved, then<br />

= + = (11.41)<br />

That is, the center-of-mass Hamiltonian equals the center-of-mass total energy. The center-of-mass<br />

Hamiltonian then can be written us<strong>in</strong>g the effective potential (1133) <strong>in</strong> the form<br />

= 2 <br />

2 +<br />

2 <br />

2 2 + () = 2 <br />

2 +<br />

2<br />

2 2 + () = 2 <br />

2 + () = (11.42)<br />

It is convenient to express the center-of-mass Hamiltonian <strong>in</strong> terms of the energy equation for the<br />

orbit <strong>in</strong> a central field us<strong>in</strong>g the transformed variable = 1 <br />

. Substitut<strong>in</strong>g equations 1133 and 1137 <strong>in</strong>to<br />

the Hamiltonian equation 1142 gives the energy equation of the orbit<br />

" µ<br />

2 # 2 <br />

+ 2 + ¡ −1¢ = (11.43)<br />

2 <br />

Energy conservation allows the Hamiltonian to be used to solve problems directly. That is, s<strong>in</strong>ce<br />

= ˙2<br />

2 + 2<br />

2 2 + () = (11.44)<br />

then<br />

s<br />

˙ = µ<br />

<br />

= ± 2<br />

− −<br />

2<br />

<br />

2 2 (11.45)<br />

The time dependence can be obta<strong>in</strong>ed by <strong>in</strong>tegration<br />

Z<br />

±<br />

= r ³<br />

´ + constant (11.46)<br />

2<br />

<br />

− −<br />

2<br />

2 2<br />

An <strong>in</strong>version of this gives the solution <strong>in</strong> the standard form = () However, it is more <strong>in</strong>terest<strong>in</strong>g to f<strong>in</strong>d<br />

the relation between and From relation 1146 for <br />

then<br />

while equation 1129 gives<br />

Therefore<br />

Z<br />

=<br />

=<br />

= <br />

2 =<br />

±<br />

r ³<br />

´ (11.47)<br />

2<br />

<br />

− −<br />

2<br />

2 2<br />

±<br />

³<br />

<br />

r2<br />

2 − −<br />

´ (11.48)<br />

2<br />

2 2<br />

±<br />

³<br />

´ + constant (11.49)<br />

<br />

r2<br />

2 − −<br />

2<br />

2 2<br />

which can be used to calculate the angular coord<strong>in</strong>ate. This gives the relation between the radial and angular<br />

coord<strong>in</strong>ates which specifies the trajectory.<br />

Although equations (1145) and (1149) formally give the solution, the actual solution can be derived<br />

analytically only for certa<strong>in</strong> specific forms of the force law and these solutions differ for attractive versus<br />

repulsive <strong>in</strong>teractions.


252 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

11.7 General features of the orbit solutions<br />

It is useful to look at the general features of the solutions of the equations of motion given by the equivalent<br />

one-body representation of the two-body motion. These orbits depend on the net center of mass energy <br />

There are five possible situations depend<strong>in</strong>g on the center-of-mass total energy .<br />

1) E 0 : The trajectory is hyperbolic and has a m<strong>in</strong>imum distance, but no maximum. The distance<br />

of closest approach is given when ˙ =0 At the turn<strong>in</strong>g po<strong>in</strong>t = +<br />

2) E = 0 : It can be shown that the orbit for this case is parabolic.<br />

3) 0 E U m<strong>in</strong> : For this case the equivalent orbit has both a maximum and m<strong>in</strong>imum radial distance<br />

at which ˙ =0 At the turn<strong>in</strong>g po<strong>in</strong>ts the radial k<strong>in</strong>etic energy term is zero so = +<br />

2<br />

2<br />

For the<br />

2<br />

attractive <strong>in</strong>verse square law force the path is an ellipse with the focus at the center of attraction (Figure<br />

115), which is Kepler’s First Law. Dur<strong>in</strong>g the time that the radius ranges from m<strong>in</strong> to max and back the<br />

radius vector turns through an angle ∆ which is given by<br />

∆ =2<br />

Z max<br />

m<strong>in</strong><br />

2<br />

2 2<br />

±<br />

³<br />

´ (11.50)<br />

<br />

r2<br />

2 − −<br />

2<br />

2 2<br />

The general path prescribes a rosette shape which is a closed curve only if ∆ is a rational fraction of<br />

2.<br />

4) E = U m<strong>in</strong> : In this case is a constant imply<strong>in</strong>g that the path is circular s<strong>in</strong>ce<br />

s<br />

˙ = <br />

= ± 2<br />

<br />

µ<br />

− −<br />

<br />

2<br />

2 2 =0 (11.51)<br />

5) E U m<strong>in</strong> : For this case the square root is imag<strong>in</strong>ary and there is no real solution.<br />

In general the orbit is not closed, and such open orbits do not repeat. Bertrand’s Theorem states that<br />

the <strong>in</strong>verse-square central force, and the l<strong>in</strong>ear harmonic oscillator, are the only radial dependences of the<br />

central force that lead to stable closed orbits.<br />

11.2 Example: Orbit equation of motion for a free body<br />

It is illustrative to use the differential orbit equation 1139 to show that<br />

a body <strong>in</strong> free motion travels <strong>in</strong> a straight l<strong>in</strong>e. Assume that a l<strong>in</strong>e through<br />

the orig<strong>in</strong> <strong>in</strong>tersects perpendicular to the <strong>in</strong>stantaneous trajectory at the<br />

po<strong>in</strong>t which has polar coord<strong>in</strong>ates ( 0 ) relative to the orig<strong>in</strong>. The<br />

po<strong>in</strong>t with polar coord<strong>in</strong>ates ( ) lies on straight l<strong>in</strong>e through that<br />

is perpendicular to if, and only if, cos( − ) = 0 S<strong>in</strong>ce the force is<br />

zero then the differential orbit equation simplifies to<br />

y<br />

P<br />

A solution of this is<br />

2 ()<br />

2 + () =0<br />

r<br />

Q<br />

() = 1 0<br />

cos( − )<br />

r 0<br />

where 0 and are arbitrary constants. This can be rewritten as<br />

x<br />

() =<br />

0<br />

cos( − )<br />

Trajectory of a free body<br />

This is the equation of a straight l<strong>in</strong>e <strong>in</strong> polar coord<strong>in</strong>ates as illustrated <strong>in</strong> the adjacent figure. This shows<br />

that a free body moves <strong>in</strong> a straight l<strong>in</strong>e if no forces are act<strong>in</strong>g on the body.


11.8. INVERSE-SQUARE, TWO-BODY, CENTRAL FORCE 253<br />

11.8 Inverse-square, two-body, central force<br />

The most important conservative, two-body, central <strong>in</strong>teraction is the attractive <strong>in</strong>verse-square law force,<br />

which is encountered <strong>in</strong> both gravitational attraction and the Coulomb force. This force F(r) can be written<br />

<strong>in</strong> the form<br />

F() = br (11.52)<br />

2 The force constant is def<strong>in</strong>ed to be negative for an attractive force and positive for a repulsive force. In<br />

S.I. units the force constant = − 1 2 for the gravitational force and =+ 1 2<br />

4 0<br />

for the Coulomb force.<br />

Note that this sign convention is the opposite of what is used <strong>in</strong> many books which use a negative sign <strong>in</strong><br />

equation 1152 and assume to be positive for an attractive force and negative for a repulsive force.<br />

The conservative, <strong>in</strong>verse-square, two-body, central force is unique <strong>in</strong> that the underly<strong>in</strong>g symmetries<br />

lead to four conservation laws, all of which are of pivotal importance <strong>in</strong> nature.<br />

1. Conservation of angular momentum: Like all conservative central forces, the <strong>in</strong>verse-square central<br />

two-body force conserves angular momentum as proven <strong>in</strong> chapter 113.<br />

2. Conservation of energy: This conservative central force can be represented <strong>in</strong> terms of a scalar<br />

potential energy () as given by equation 112 where for this central force<br />

() = <br />

(11.53)<br />

Moreover, equation 1142 showed that the center-of-mass Hamiltonian is conserved, that is, = <br />

3. Gauss’ Law: For a conservative, <strong>in</strong>verse-square, two-body, central force, the flux of the force field out<br />

of any closed surface is proportional to the algebraic sum of the sources and s<strong>in</strong>ks of this field that<br />

are located <strong>in</strong>side the closed surface. The net flux is <strong>in</strong>dependent of the distribution of the sources<br />

and s<strong>in</strong>ks <strong>in</strong>side the closed surface, as well as the size and shape of the closed surface. Chapter 2145<br />

proved this for the gravitational force field.<br />

4. Closed orbits: Two bodies <strong>in</strong>teract<strong>in</strong>g via the conservative, <strong>in</strong>verse-square, two-body, central force<br />

follow closed (degenerate) orbits as stated by Bertrand’s Theorem. The first consequence of this<br />

symmetry is that Kepler’s laws of planetary motion have stable, s<strong>in</strong>gle-valued orbits. The second<br />

consequence of this symmetry is the conservation of the eccentricity vector discussed <strong>in</strong> chapter 1184.<br />

Observables that depend on Gauss’s Law, or on closed planetary orbits, are extremely sensitive to addition<br />

of even a m<strong>in</strong>iscule <strong>in</strong>cremental exponent to the radial dependence −(2±) of the force. The statement<br />

that the <strong>in</strong>verse-square, two-body, central force leads to closed orbits can be proven by <strong>in</strong>sert<strong>in</strong>g equation<br />

1152 <strong>in</strong>to the orbit differential equation,<br />

2 <br />

2 + = − 1<br />

2 2 2 = − <br />

2 (11.54)<br />

Us<strong>in</strong>g the transformation<br />

≡ + <br />

2 (11.55)<br />

the orbit equation becomes<br />

2 <br />

2<br />

+ =0 (11.56)<br />

<br />

A solution of this equation is<br />

= cos ( − 0 ) (11.57)<br />

Therefore<br />

= 1 = − 2 [1 + cos ( − 0)] (11.58)<br />

This the equation of a conic section. For an attractive, <strong>in</strong>verse-square, central force, equation 1158 is the<br />

equation for an ellipse with the orig<strong>in</strong> of at one of the foci of the ellipse that has eccentricity def<strong>in</strong>ed as<br />

≡ 2<br />

<br />

(11.59)


254 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

Equation 1158 is the polar equation of a conic section. Equation 1158 also can be derived with the<br />

orig<strong>in</strong> at a focus by <strong>in</strong>sert<strong>in</strong>g the <strong>in</strong>verse square law potential <strong>in</strong>to equation 1149 which gives<br />

Z<br />

±<br />

= q<br />

+ constant (11.60)<br />

2 <br />

<br />

+ 2<br />

2 <br />

− 2 2<br />

The solution of this gives<br />

" s<br />

#<br />

= 1 = − 2 1+ 1+ 2 2<br />

2 cos ( − 0 )<br />

(11.61)<br />

Equations 1158 and 1161 are identical if the eccentricity equals<br />

s<br />

=<br />

1+ 2 2<br />

2 (11.62)<br />

The value of 0 merely determ<strong>in</strong>es the orientation of the major axis of the equivalent orbit. Without loss of<br />

generality, it is possible to assume that the angle is measured with respect to the major axis of the orbit,<br />

that is 0 =0. Then the equation can be written as<br />

" s<br />

#<br />

= 1 = − [1 + cos ()] = −<br />

2 2 1+ 1+ 2 2<br />

2 cos ()<br />

(11.63)<br />

This is the equation of a conic section where is the eccentricity of the conic section. The conic section is a<br />

hyperbola if 1, parabola if =1 ellipse if 1 and a circle if =0 All the equivalent one-body orbits<br />

for an attractive force have the orig<strong>in</strong> of the force at a focus of the conic section. The orbits depend on<br />

whether the force is attractive or repulsive, on the conserved angular momentum andonthecenter-of-mass<br />

energy .<br />

11.8.1 Bound orbits<br />

Closed bound orbits occur only if the follow<strong>in</strong>g requirements<br />

are satisfied.<br />

1. The force must be attractive, ( 0) then equation<br />

1163 ensures that is positive.<br />

2. For a closed elliptical orbit. the eccentricity 1 of the<br />

equivalent one-body representation of the orbit implies<br />

that the total center-of-mass energy 0, thatis,<br />

the closed orbit is bound.<br />

Bound elliptical orbits have the center-of-force at one <strong>in</strong>terior<br />

focus 1 of the elliptical one-body representation of the<br />

orbit as shown <strong>in</strong> figure 115.<br />

The m<strong>in</strong>imum value of the orbit = m<strong>in</strong> occurs when<br />

=0 where<br />

m<strong>in</strong> = −<br />

[1 + ]<br />

This m<strong>in</strong>imum distance is called the periapsis 1 .<br />

2<br />

(11.64)<br />

Figure 11.5: Bound elliptical orbit.<br />

1 The greek term apsis refers to the po<strong>in</strong>ts of greatest or least distance of approach for an orbit<strong>in</strong>g body from one of the<br />

foci of the elliptical orbit. The term periapsis or pericenter both are used to designate the closest distance of approach, while<br />

apoapsis or apocenter are used to designate the farthest distance of approach. Attach<strong>in</strong>g the terms "perí-" and "apo-" to the<br />

general term "-apsis" is preferred over hav<strong>in</strong>g different names for each object <strong>in</strong> the solar system. For example, frequently used<br />

terms are "-helion" for orbits of the sun, "-gee" for orbits around the earth, and "-cynthion" for orbits around the moon.


11.8. INVERSE-SQUARE, TWO-BODY, CENTRAL FORCE 255<br />

The maximum distance, = max which is called the apoapsis, occurs when = 180 <br />

2<br />

max = −<br />

[1 − ]<br />

(11.65)<br />

Remember that s<strong>in</strong>ce 0 for bound orbits, the negative signs <strong>in</strong> equations 1164 and 1165 lead to 0.<br />

The most bound orbit is a circle hav<strong>in</strong>g =0which implies that = − 2<br />

<br />

. 2<br />

The shape of the elliptical orbit also can be described with respect to the center of the elliptical equivalent<br />

orbit by deriv<strong>in</strong>g the lengths of the semi-major axis and the semi-m<strong>in</strong>or axis shown <strong>in</strong> figure 115<br />

= 1 2 ( m<strong>in</strong> + max )= 1 µ<br />

2<br />

2 [1 + ] + 2 <br />

2<br />

=<br />

[1 − ] [1 − 2 (11.66)<br />

]<br />

= p 1 − 2 =<br />

p [1 − 2 ]<br />

2<br />

(11.67)<br />

Remember that the predicted bound elliptical orbit corresponds to the equivalent one-body representation<br />

for the two-body motion as illustrated <strong>in</strong> figure 112. This can be transformed to the <strong>in</strong>dividual spatial<br />

trajectories of the each of the two bodies <strong>in</strong> an <strong>in</strong>ertial frame.<br />

11.8.2 Kepler’s laws for bound planetary motion<br />

Kepler’s three laws of motion apply to the motion of two bodies <strong>in</strong> a bound orbit due to the attractive<br />

gravitational force for which = − 1 2 .<br />

1) Each planet moves <strong>in</strong> an elliptical orbit with the sun at one focus<br />

2) The radius vector, drawn from the sun to a planet, describes equal areas <strong>in</strong> equal times<br />

3) The square of the period of revolution about the sun is proportional to the cube of the major axis<br />

of the orbit.<br />

Two bodies <strong>in</strong>teract<strong>in</strong>g via the gravitational force, which is a conservative, <strong>in</strong>verse-square, two-body<br />

central force, is best handled us<strong>in</strong>g the equivalent orbit representation. The first and second laws were<br />

proved <strong>in</strong> chapters 118 and 113. That is, the second law is equivalent to the statement that the angular<br />

momentum is conserved. The third law can be derived us<strong>in</strong>g the fact that the area of an ellipse is<br />

= = 2p 1 − 2 =<br />

√ −<br />

3 2 (11.68)<br />

Equations 1126 and 1127 give that the rate of change of area swept out by the radius vector is<br />

<br />

= 1 <br />

2 2 ˙ =<br />

2<br />

Therefore the period for one revolution is given by the time to sweep out one complete ellipse<br />

(11.69)<br />

=<br />

<br />

¡ <br />

<br />

µ 1<br />

<br />

2<br />

3<br />

¢ =2 2 (11.70)<br />

−<br />

This leads to Kepler’s 3 law<br />

2 =4 2 − 3 (11.71)<br />

Bound orbits occur only for attractive forces for which the force constant is negative, and thus cancel<br />

the negative sign <strong>in</strong> equation 1171. For example, for the gravitational force = − 1 2 .<br />

Note that the reduced mass =<br />

12<br />

1 + 2<br />

occurs <strong>in</strong> Kepler’s 3 law. That is, Kepler’s third law can be<br />

written <strong>in</strong> terms of the actual masses of the bodies to be<br />

2 =<br />

4 2<br />

( 1 + 2 ) 3 (11.72)<br />

In relat<strong>in</strong>g the relative periods of the different planets Kepler made the approximation that the mass of the<br />

planet 1 is negligible relative to the mass of the sun 2


256 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

The eccentricity of the major planets ranges from =02056 for Mercury, to =00068 for Venus. The<br />

Earth has an eccentricity of =00167 with m<strong>in</strong> =91· 10 6 miles and max =95· 10 6 miles. On the other<br />

hand, =0967 for Halley’s comet, that is, the radius vector ranges from 06 to 18 times the radius of the<br />

orbit of the Earth.<br />

The orbit energy can be derived by substitut<strong>in</strong>g the eccentricity, given by equation 1162 <strong>in</strong>to the semimajor<br />

axis length given by equation 1166 which leads to the center-of-mass energy of<br />

= − 2<br />

However, the Hamiltonian, given by equation 1142 implies that is<br />

= 1 µ<br />

2 2 + − <br />

= − 2<br />

(11.73)<br />

(11.74)<br />

For the simple case of a circular orbit, = then the velocity equals<br />

s<br />

<br />

=<br />

(11.75)<br />

<br />

For a circular orbit, the drag on a satellite lowers the total energy result<strong>in</strong>g <strong>in</strong> a decrease <strong>in</strong> the radius<br />

of the orbit and a concomitant <strong>in</strong>crease <strong>in</strong> velocity. That is, when the orbit radius is decreased, part of the<br />

ga<strong>in</strong> <strong>in</strong> potential energy accounts for the work done aga<strong>in</strong>st the drag, and the rema<strong>in</strong><strong>in</strong>g part goes towards<br />

<strong>in</strong>crease of the k<strong>in</strong>etic energy. Also note that, as predicted by the Virial Theorem, the k<strong>in</strong>etic energy always<br />

is half the potential energy for the <strong>in</strong>verse square law force.<br />

11.8.3 Unbound orbits<br />

Attractive <strong>in</strong>verse-square central forces lead to hyperbolic<br />

orbits for 1 for which 0, that is, the orbit is<br />

unbound. In addition, the orbits always are unbound for<br />

arepulsiveforces<strong>in</strong>ce = <br />

is positive as is the k<strong>in</strong>etic<br />

energy ,thus = + 0 The radial orbit<br />

equation for either an attractive or a repulsive force is<br />

= −<br />

(11.76)<br />

[1 + cos ]<br />

For a repulsive force is positive and 2 always is positive.<br />

Therefore to ensure that rema<strong>in</strong> positive the bracket term<br />

must be negative. That is<br />

2<br />

[1 + cos ] 0 0 (11.77)<br />

For an attractive force is negative and s<strong>in</strong>ce 2 is positive<br />

then the bracket term must be positive to ensure that is<br />

positive. That is,<br />

[1 + cos ] 0 0 (11.78)<br />

Figure 116 shows both branches of the hyperbola for a given<br />

angle for the equivalent two-body orbits where the center<br />

of force is at the orig<strong>in</strong>. For an attractive force, 0<br />

the center of force is at the <strong>in</strong>terior focus of the hyperbola,<br />

whereas for a repulsive force the center of force is at the<br />

exterior focus. For a given value of || the asymptotes of the<br />

orbits both are displaced by the same impact parameter<br />

from parallel l<strong>in</strong>es pass<strong>in</strong>g through the center of force.<br />

The scatter<strong>in</strong>g angle, between the outgo<strong>in</strong>g direction of the<br />

scattered body and the <strong>in</strong>cident direction, is designated to<br />

be which is related to the angle by = 180 ◦ − 2.<br />

Figure 11.6: Hyperbolic two-body orbits for a<br />

repulsive (left) and attractive (right) <strong>in</strong>versesquare,<br />

central two-body forces. Both orbits<br />

have the angular momentum vector po<strong>in</strong>t<strong>in</strong>g<br />

upwards out of the plane of the orbit


11.8. INVERSE-SQUARE, TWO-BODY, CENTRAL FORCE 257<br />

11.8.4 Eccentricity vector<br />

Two-bodies <strong>in</strong>teract<strong>in</strong>g via a conservative two-body central force have two <strong>in</strong>variant first-order <strong>in</strong>tegrals,<br />

namely the conservation of energy and the conservation of angular momentum. For the special case of the<br />

<strong>in</strong>verse-square law, there is a third <strong>in</strong>variant of the motion, which Hamilton called the eccentricity vector 2 ,<br />

that unambiguously def<strong>in</strong>es the orientation and direction of the major axis of the elliptical orbit. It will be<br />

shown that the angular momentum plus the eccentricity vector completely def<strong>in</strong>e the plane and orientation<br />

of the orbit for a conservative <strong>in</strong>verse-square law central force.<br />

Newton’s second law for a central force can be written <strong>in</strong> the form<br />

ṗ =()ˆr (11.79)<br />

Note that the angular moment L = r × p is conserved for a central force, that is ˙L =0. Therefore the time<br />

derivative of the product p × L reduces to<br />

<br />

(p × L)=ṗ × L =()ˆr× (r×ṙ) =() £ ¤ r (r · ṙ) − 2ṙ <br />

(11.80)<br />

This can be simplified us<strong>in</strong>g the fact that<br />

r · ṙ = 1 <br />

(r · r) = ˙<br />

2 <br />

(11.81)<br />

thus<br />

() £ ¤ ∙ṙ<br />

r (r · ṙ) − 2ṙ = −() 2<br />

<br />

− r ˙ ¸<br />

2 = −() 2 ³ r<br />

´<br />

<br />

(11.82)<br />

This allows equation 1180 to be reduced to<br />

<br />

(p × L)=− ³ r<br />

´<br />

()2 <br />

(11.83)<br />

Assume the special case of the <strong>in</strong>verse-square law, equation 1152, then the central force equation 1183<br />

reduces to<br />

<br />

(p × L)=− (ˆr)<br />

<br />

(11.84)<br />

or<br />

Def<strong>in</strong>e the eccentricity vector A as<br />

<br />

[(p × L)+(ˆr)] = 0 (11.85)<br />

<br />

A ≡ (p × L)+(ˆr) (11.86)<br />

then equation 1185 corresponds to<br />

A<br />

=0 (11.87)<br />

<br />

This is a statement that the eccentricity vector is a constant of motion for an <strong>in</strong>verse-square, central<br />

force.<br />

The def<strong>in</strong>ition of the eccentricity vector A and angular momentum vector L implies a zero scalar product,<br />

A · L =0 (11.88)<br />

Thus the eccentricity vector A and angular momentum L are mutually perpendicular, that is, A is <strong>in</strong> the<br />

plane of the orbit while L is perpendicular to the plane of the orbit. The eccentricity vector A, always po<strong>in</strong>ts<br />

along the major axis of the ellipse from the focus to the periapsis as illustrated on the left side <strong>in</strong> figure 117.<br />

2 The symmetry underly<strong>in</strong>g the eccentricity vector is less <strong>in</strong>tuitive than the energy or angular momentum <strong>in</strong>variants lead<strong>in</strong>g<br />

to it be<strong>in</strong>g discovered <strong>in</strong>dependently several times dur<strong>in</strong>g the past three centuries. Jakob Hermann was the first to <strong>in</strong>dentify<br />

this <strong>in</strong>variant for the special case of the <strong>in</strong>verse-square central force. Bernoulli generalized his proof <strong>in</strong> 1710. Laplace derived<br />

the <strong>in</strong>variant at the end of the 18 century us<strong>in</strong>g analytical mechanics. Hamilton derived the connection between the <strong>in</strong>variant<br />

and the orbit eccentricity. Gibbs derived the <strong>in</strong>variant us<strong>in</strong>g vector analysis. Runge published the Gibb’s derivation <strong>in</strong> his<br />

textbook which was referenced by Lenz <strong>in</strong> a 1924 paper on the quantal model of the hydrogen atom. Goldste<strong>in</strong> named this<br />

<strong>in</strong>variant the "Laplace-Runge-Lenz vector", while others have named it the "Runge-Lenz vector" or the "Lenz vector". This<br />

book uses Hamilton’s more <strong>in</strong>tuitive name of "eccentricity vector".


258 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

Figure 11.7: The elliptical trajectory and eccentricity vector A for two bodies <strong>in</strong>teract<strong>in</strong>g via the <strong>in</strong>versesquare,<br />

central force for eccentricity =075. The left plot shows the elliptical spatial trajectory where<br />

the semi-major axis is assumed to be on the -axis and the angular momentum L =ẑ, is out of the page.<br />

The force centre is at one foci of the ellipse. The vector coupl<strong>in</strong>g relation A ≡ (p × L)+(ˆr) is illustrated<br />

at four po<strong>in</strong>ts on the spatial trajectory. The right plot is a hodograph of the l<strong>in</strong>ear momentum p for this<br />

trajectory. The periapsis is denoted by the number 1 and the apoapsis is marked as 3 on both plots. Note<br />

that the eccentricity vector A is a constant that po<strong>in</strong>ts parallel to the major axis towards the perapsis.<br />

As a consequence, the two orthogonal vectors A and L completely def<strong>in</strong>e the plane of the orbit, plus the<br />

orientation of the major axis of the Kepler orbit, <strong>in</strong> this plane. The three vectors A, p × L, and(ˆr) obey<br />

thetriangleruleasillustrated<strong>in</strong>theleftsideoffigure 117.<br />

Hamilton noted the direct connection between the eccentricity vector A and the eccentricity of the<br />

conic section orbit. This can be shown by consider<strong>in</strong>g the scalar product<br />

Notethatthetriplescalarproductcanbepermutedtogive<br />

A · r = cos = r· (p × L)+ (11.89)<br />

r· (p × L) =(r × p) ·L = L · L = 2 (11.90)<br />

Insert<strong>in</strong>g equation 1190 <strong>in</strong>to 1189 gives<br />

µ<br />

1<br />

= − 2 1 − <br />

cos <br />

(11.91)<br />

Note that equations 1163 and 1191 are identical if 0 =0. This implies that the eccentricity and <br />

are related by<br />

= − (11.92)<br />

<br />

where is def<strong>in</strong>ed to be negative for an attractive force. The relation between the eccentricity and total<br />

center-of-mass energy can be used to rewrite equation 1162 <strong>in</strong> the form<br />

2 = 2 2 +2 2 (11.93)<br />

The comb<strong>in</strong>ation of the eccentricity vector A and the angular momentum vector L completely specifies<br />

the orbit for an <strong>in</strong>verse square-law central force. The trajectory is <strong>in</strong> the plane perpendicular to the angular<br />

momentum vector L, while the eccentricity, plus the orientation of the orbit, both are def<strong>in</strong>ed by the<br />

eccentricity vector A. The eccentricity vector and angular momentum vector each have three <strong>in</strong>dependent<br />

coord<strong>in</strong>ates, that is, these two vector <strong>in</strong>variants provide six constra<strong>in</strong>ts, while the scalar <strong>in</strong>variant energy <br />

adds one additional constra<strong>in</strong>t. The exact location of the particle mov<strong>in</strong>g along the trajectory is not def<strong>in</strong>ed<br />

and thus there are only five <strong>in</strong>dependent coord<strong>in</strong>ates governed by the above seven constra<strong>in</strong>ts. Thus the


11.9. ISOTROPIC, LINEAR, TWO-BODY, CENTRAL FORCE 259<br />

eccentricity vector, angular momentum, and center-of-mass energy are related by the two equations 1188<br />

and 1193.<br />

Noether’s theorem states that each conservation law is a manifestation of an underly<strong>in</strong>g symmetry.<br />

Identification of the underly<strong>in</strong>g symmetry responsible for the conservation of the eccentricity vector A is<br />

elucidated us<strong>in</strong>g equation 1186 to give<br />

(ˆr) =A− (p × L) (11.94)<br />

Take the scalar product<br />

(ˆr) · (ˆr) =() 2 = 2 2 + 2 − 2 · (p × L) (11.95)<br />

Choose the angular momentum to be along the -axis, that is, L =ẑ, and, s<strong>in</strong>ce p and A are perpendicular<br />

to L, thenp and A are <strong>in</strong> the ˆx − ŷ plane. Assume that the semimajor axis of the elliptical orbit is along<br />

the x-axis, then the locus of the momentum vector on a momentum hodograph has the equation<br />

µ<br />

2 + − 2<br />

=<br />

<br />

µ 2 <br />

(11.96)<br />

<br />

¯<br />

Equation 1196 implies that the locus of the momentum vector is a circle of radius<br />

¯<br />

¯ <br />

<br />

¯ with the center<br />

displaced from the orig<strong>in</strong> at coord<strong>in</strong>ates ¡ 0 <br />

¢<br />

as shown by the momentum hodograph on the right side of<br />

an figure 117. The angle and eccentricity are related by,<br />

cos = − <br />

= − = (11.97)<br />

The circular orbit is centered at the orig<strong>in</strong> for = − <br />

=0, and thus the magnitude |p| is a constant around<br />

the whole trajectory.<br />

The <strong>in</strong>verse-square, central, two-body, force is unusual <strong>in</strong> that it leads to stable closed bound orbits<br />

because the radial and angular frequencies are degenerate, i.e. = In momentum space, the locus of<br />

the l<strong>in</strong>ear momentum vector p is a perfect circle which is the underly<strong>in</strong>g symmetry responsible for both the<br />

fact that the orbits are closed, and the <strong>in</strong>variance of the eccentricity vector. Mathematically this symmetry<br />

for the Kepler problem corresponds to the body mov<strong>in</strong>g freely on the boundary of a four-dimensional sphere<br />

<strong>in</strong> space and momentum. The <strong>in</strong>variance of the eccentricity vector is a manifestation of the special property<br />

of the <strong>in</strong>verse-square, central force under certa<strong>in</strong> rotations <strong>in</strong> this four-dimensional space; this (4) symmetry<br />

is an example of a hidden symmetry.<br />

11.9 Isotropic, l<strong>in</strong>ear, two-body, central force<br />

Closed orbits occur for the two-dimensional l<strong>in</strong>ear oscillator when <br />

<br />

is a rational fraction as discussed <strong>in</strong><br />

chapter 33. Bertrand’s Theorem states that the l<strong>in</strong>ear oscillator, and the <strong>in</strong>verse-square law (Kepler<br />

problem), are the only two-body central forces that have s<strong>in</strong>gle-valued, stable, closed orbits of the coupled<br />

radial and angular motion. The <strong>in</strong>variance of the eccentricity vector was the underly<strong>in</strong>g symmetry lead<strong>in</strong>g<br />

to s<strong>in</strong>gle-valued, stable, closed orbits for the Kepler problem. It is <strong>in</strong>terest<strong>in</strong>g to explore the symmetry that<br />

leads to stable closed orbits for the harmonic oscillator. For simplicity, this discussion will restrict discussion<br />

to the isotropic, harmonic, two-body, central force where = = , for which the two-body, central force<br />

is l<strong>in</strong>ear<br />

F() =r (11.98)<br />

where 0 corresponds to a repulsive force and 0 to an attractive force. This isotropic harmonic force<br />

can be expressed <strong>in</strong> terms of a spherical potential () where<br />

() =− 1 2 2 (11.99)<br />

S<strong>in</strong>ce this is a central two-body force, both the equivalent one-body representation, and the conservation<br />

of angular momentum, are equally applicable to the harmonic two-body force. As discussed <strong>in</strong> section<br />

113, s<strong>in</strong>ce the two-body force is central, the motion is conf<strong>in</strong>ed to a plane, and thus the Lagrangian can


260 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

be expressed <strong>in</strong> polar coord<strong>in</strong>ates. In addition, s<strong>in</strong>ce the force is spherically symmetric, then the angular<br />

momentum is conserved. The orbit solutions are conic sections as described <strong>in</strong> chapter 117. The shape of<br />

the orbit for the harmonic two-body central force can be derived us<strong>in</strong>g either polar or cartesian coord<strong>in</strong>ates<br />

as illustrated below.<br />

11.9.1 Polar coord<strong>in</strong>ates<br />

The orig<strong>in</strong> of the equivalent orbit for the harmonic force will be found to be at the center of an ellipse, rather<br />

thanthefocioftheellipseasfoundforthe<strong>in</strong>versesquarelaw.Theshapeoftheorbitcanbedef<strong>in</strong>ed us<strong>in</strong>g<br />

aB<strong>in</strong>etdifferential orbit equation that employs the transformation<br />

Then<br />

0 ≡ 1 2 (11.100)<br />

0<br />

= − 2 3 <br />

<br />

The cha<strong>in</strong> rule gives that<br />

˙ = <br />

˙ = − 3 0 ˙<br />

2 = − 0<br />

2 <br />

Substitute this <strong>in</strong>to the Hamiltonian equation 1142 gives<br />

1<br />

2 ˙2 = 1 2 µ<br />

<br />

0<br />

8 0 <br />

Rearrang<strong>in</strong>g this equation gives<br />

µ <br />

0 2<br />

+4 02 − 8<br />

<br />

2 0 = 4<br />

<br />

2 <br />

Addition of a constant to both sides of the equation completes the square<br />

" Ã !# 2 ! 2 Ã<br />

<br />

0 − +4Ã<br />

<br />

2 0 − <br />

<br />

2 =+ 4<br />

<br />

2 +4<br />

<br />

(11.101)<br />

(11.102)<br />

2<br />

= − 2 <br />

2 0 + <br />

2 0 (11.103)<br />

(11.104)<br />

! 2<br />

<br />

2 (11.105)<br />

<br />

The right-hand side of equation 11105 is a constant. The solution of 11105 must be a s<strong>in</strong>e or cos<strong>in</strong>e function<br />

with polar angle = . Thatis<br />

à ! ⎡Ã<br />

! ⎤ 1<br />

2<br />

0 − <br />

2 = ⎣<br />

<br />

<br />

2 + <br />

2<br />

⎦<br />

<br />

2 cos 2 ( − 0 ) (11.106)<br />

<br />

That is,<br />

0 = 1 2 = <br />

2 <br />

⎛ Ã ! 1<br />

⎝1+ 1+ 2 2<br />

<br />

2 cos 2( − <br />

<br />

0 )⎞<br />

⎠ (11.107)<br />

Equation 11107 corresponds to a closed orbit centered at the orig<strong>in</strong> of the elliptical orbit as illustrated <strong>in</strong><br />

figure 118 The eccentricity of this closed orbit is given by<br />

à ! 1<br />

1+ 2 2<br />

<br />

2 = 2<br />

2 − 2 (11.108)<br />

Equations 1166 1167 give that the eccentricity is related to the semi-major and semi-m<strong>in</strong>or axes by<br />

µ 2 <br />

2 =1−<br />

(11.109)<br />

<br />

Note that for a repulsive force 0, then ≥ 1 lead<strong>in</strong>g to unbound hyperbolic or parabolic orbits centered<br />

on the orig<strong>in</strong>. An attractive force, 0 allows for bound elliptical, as well as unbound parabolic and<br />

hyperbolic orbits.


11.9. ISOTROPIC, LINEAR, TWO-BODY, CENTRAL FORCE 261<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

-1.2 -1.0 -0.8 -0.6 -0.4 -0.2 0.2 0.4 0.6 0.8 1.0 1.2<br />

-0.2<br />

-0.4<br />

-0.6<br />

-0.8<br />

y<br />

r<br />

x<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

p<br />

y<br />

0.4<br />

0.2<br />

p x<br />

-0.8 -0.6 -0.4 -0.2 0.2 0.4 0.6 0.8<br />

-0.2<br />

-0.4<br />

-0.6<br />

-0.8<br />

-1.0<br />

-1.2<br />

p<br />

Figure 11.8: The elliptical equivalent trajectory for two bodies <strong>in</strong>teract<strong>in</strong>g via the l<strong>in</strong>ear, central force for<br />

eccentricity =075. The left plot shows the elliptical spatial trajectory where the semi-major axis is<br />

assumedtobeonthe-axis and the angular momentum L =ẑ, is out of the page. The force center is at<br />

the center of the ellipse. The right plot is a hodograph of the l<strong>in</strong>ear momentum p for this trajectory.<br />

11.9.2 Cartesian coord<strong>in</strong>ates<br />

The isotropic harmonic oscillator, expressed <strong>in</strong> terms of cartesian coord<strong>in</strong>ates <strong>in</strong> the ( ) plane of the orbit,<br />

is separable because there is no direct coupl<strong>in</strong>g term between the and motion. That is. the center-of-mass<br />

Lagrangian <strong>in</strong> the ( ) plane separates <strong>in</strong>to <strong>in</strong>dependent motion for and .<br />

= 1 2 ṙ · ṙ + 1 ∙ 1<br />

2 r · r = 2 ˙2 + 1 2¸ ∙ 1<br />

+<br />

2 2 ˙2 + 1 2¸<br />

(11.110)<br />

2<br />

Solutions for the <strong>in</strong>dependent coord<strong>in</strong>ates, and their correspond<strong>in</strong>g momenta, are<br />

where =<br />

q<br />

<br />

. Therefore<br />

r = ˆ cos ( + )+ˆ cos ( + ) (11.111)<br />

p = −ˆ s<strong>in</strong> ( + ) − ˆ s<strong>in</strong> ( + ) (11.112)<br />

where<br />

2 = 2 + 2 =[ cos ( + )] 2 +[ cos ( + )] 2 (11.113)<br />

p<br />

= 2 + 2 4 + <br />

+<br />

4 +2 2 cos ( − )<br />

cos (2 + <br />

2<br />

2<br />

0 )<br />

cos 0 =<br />

2 cos + 2 cos <br />

p<br />

4 + 4 +2 2 cos ( − )<br />

(11.114)<br />

For a phase difference − = ± 2<br />

this equation describes an ellipse centered at the orig<strong>in</strong> which agrees<br />

with equation 11107 that was derived us<strong>in</strong>g polar coord<strong>in</strong>ates.<br />

The two normal modes of the isotropic harmonic oscillator are degenerate, therefore are equally good<br />

normal modes with two correspond<strong>in</strong>g total energies, 1 2 , while the correspond<strong>in</strong>g angular momentum <br />

po<strong>in</strong>ts <strong>in</strong> the direction.<br />

1 = 2 <br />

2 + 1 2 2 (11.115)<br />

2 = 2 <br />

2 + 1 2 2 (11.116)<br />

= ( − ) (11.117)


262 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

Figure 118 shows the closed elliptical equivalent orbit plus the correspond<strong>in</strong>g momentum hodograph for<br />

the isotropic harmonic two-body central force. Figures 117 and 118 contrast the differences between the<br />

elliptical orbits for the <strong>in</strong>verse-square force, and those for the harmonic two-body central force. Although<br />

the orbits for bound systems with the harmonic two-body force, and the <strong>in</strong>verse-square force, both lead to<br />

elliptical bound orbits, there are important differences. Both the radial motion and momentum are two<br />

valuedpercycleforthereflection-symmetric harmonic oscillator, whereas the radius and momentum have<br />

only one maximum and one m<strong>in</strong>imum per revolution for the <strong>in</strong>verse-square law. Although the <strong>in</strong>verse-square,<br />

and the isotropic, harmonic, two-body central forces both lead to closed bound elliptical orbits for which the<br />

angular momentum is conserved and the orbits are planar, there is another important difference between the<br />

orbits for these two <strong>in</strong>teractions. The orbit equation for the Kepler problem is expressed with respect to a<br />

foci of the elliptical equivalent orbit, as illustrated <strong>in</strong> figure 117, whereas the orbit equation for the isotropic<br />

harmonic oscillator orbit is expressed with respect to the center of the ellipse as illustrated <strong>in</strong> figure 118.<br />

11.9.3 Symmetry tensor A 0<br />

The <strong>in</strong>variant vectors L and A provide a complete specification of the geometry of the bound orbits for<br />

the <strong>in</strong>verse square-law Kepler system. It is <strong>in</strong>terest<strong>in</strong>g to search for a similar <strong>in</strong>variant that fully specifies<br />

the orbits for the isotropic harmonic central force. In contrast to the Kepler problem, the harmonic force<br />

center is at the center of the elliptical orbit, and the orbit is reflection symmetric with the radial and angular<br />

frequencies related by =2 . S<strong>in</strong>ce the orbit is reflection-symmetric, the orientation of the major axis<br />

of the orbit cannot be uniquely specified by a vector. Therefore, for the harmonic <strong>in</strong>teraction it is necessary<br />

to specify the orientation of the pr<strong>in</strong>cipal axis by the symmetry tensor. The symmetry of the isotropic<br />

harmonic, two-body, central force leads to the symmetry tensor A 0 which is an <strong>in</strong>variant of the motion<br />

analogous to the eccentricity vector A. Like a rotation matrix, the symmetry tensor def<strong>in</strong>es the orientation,<br />

but not direction, of the major pr<strong>in</strong>cipal axis of the elliptical orbit. In the plane of the polar orbit the 3 × 3<br />

symmetry tensor A 0 reduces to a 2 × 2 matrix hav<strong>in</strong>g matrix elements def<strong>in</strong>ed to be,<br />

0 = <br />

2 + 1 2 (11.118)<br />

The diagonal matrix elements 0 11 = 1,and 0 22 = 2 which are constants of motion. The off-diagonal<br />

term is given by<br />

µ<br />

02 <br />

12 ≡<br />

2 + 1 2 µ <br />

2<br />

2 = <br />

2 + 1 Ã !<br />

2<br />

2 2 <br />

2 + 1 2 2 − 4 ( − ) 2 = 1 2 − 2<br />

4 3 (11.119)<br />

The terms on the right-hand side of equation 11119 all are constants of motion, therefore 02<br />

12 also is a<br />

constant of motion. Thus the 3×3 symmetry tensor A 0 can be reduced to a 2×2 symmetry tensor for which<br />

all the matrix elements are constants of motion, and the trace of the symmetry tensor is equal to the total<br />

energy.<br />

In summary, the <strong>in</strong>verse-square, and harmonic oscillator two-body central <strong>in</strong>teractions both lead to closed,<br />

elliptical equivalent orbits, the plane of which is perpendicular to the conserved angular momentum vector.<br />

However, for the <strong>in</strong>verse-square force, the orig<strong>in</strong> of the equivalent orbit is at the focus of the ellipse and<br />

= , whereas the orig<strong>in</strong> is at the center of the ellipse and =2 for the harmonic force. As a<br />

consequence, the elliptical orbit is reflection symmetric for the harmonic force but not for the <strong>in</strong>verse square<br />

force. The eccentricity vector and symmetry tensor both specify the major axes of these elliptical orbits,<br />

the plane of which are perpendicular to the angular momentum vector. The eccentricity vector, and the<br />

symmetry tensor, both are directly related to the eccentricity of the orbit and the total energy of the twobody<br />

system. Noether’s theorem states that the <strong>in</strong>variance of the eccentricity vector and symmetry tensor,<br />

plus the correspond<strong>in</strong>g closed orbits, are manifestations of underly<strong>in</strong>g symmetries. The dynamical 3<br />

symmetry underlies the <strong>in</strong>variance of the symmetry tensor, whereas the dynamical 4 symmetry underlies<br />

the <strong>in</strong>variance of the eccentricity vector. These symmetries lead to stable closed elliptical bound orbits only<br />

for these two specific two-body central forces, and not for other two-body central forces.


11.10. CLOSED-ORBIT STABILITY 263<br />

11.10 Closed-orbit stability<br />

Bertrand’s theorem states that the l<strong>in</strong>ear oscillator and<br />

the <strong>in</strong>verse-square law are the only two-body, central<br />

forces for which all bound orbits are s<strong>in</strong>gle-valued, and<br />

stable closed orbits. The stability of closed orbits can<br />

be illustrated by study<strong>in</strong>g their response to perturbations.<br />

For simplicity, the follow<strong>in</strong>g discussion of stability<br />

will focus on circular orbits, but the general pr<strong>in</strong>ciples<br />

are the same for elliptical orbits.<br />

A circular orbit occurs whenever the attractive<br />

force just balances the effective ”centrifugal force” <strong>in</strong><br />

the rotat<strong>in</strong>g frame. This can occur for any radial functional<br />

form for the central force. The effective potential,<br />

equation 1133 will have a stationary po<strong>in</strong>t when<br />

µ <br />

<br />

=0 (11.120)<br />

<br />

= 0<br />

that is, when<br />

µ <br />

−<br />

<br />

= 2<br />

0<br />

0<br />

3 =0 (11.121)<br />

This is equivalent to the statement that the net force<br />

is zero. S<strong>in</strong>ce the central attractive force is given by<br />

() =− <br />

<br />

then the stationary po<strong>in</strong>t occurs when<br />

( 0 )=− 2<br />

3 0<br />

= − 0 ˙2<br />

(11.122)<br />

(11.123)<br />

This is the so-called centrifugal force <strong>in</strong> the rotat<strong>in</strong>g<br />

frame. The Hamiltonian, equation 1144, givesthat<br />

s µ<br />

<br />

2<br />

˙ = ± − −<br />

2<br />

<br />

2 2 (11.124)<br />

For a circular orbit ˙ =0that is<br />

= −<br />

2<br />

2 2 (11.125)<br />

A stable circular orbit is possible if both equations<br />

(11121) and (11125) are satisfied. Such a circular<br />

orbit will be a stable orbit at the m<strong>in</strong>imum when<br />

µ 2 <br />

<br />

2 0 (11.126)<br />

= 0<br />

Figure 11.9: Stable and unstable effective central<br />

potentials. The repulsive centrifugal and the attractive<br />

potentials (k


264 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

which can be written <strong>in</strong> terms of the central force for a stable orbit as<br />

µ <br />

− + 3 ( 0)<br />

0 (11.128)<br />

<br />

0<br />

0<br />

If the attractive central force can be expressed as a power law<br />

() =− (11.129)<br />

then stability requires<br />

0 −1 (3 + ) 0 (11.130)<br />

or<br />

−3 (11.131)<br />

Stable equivalent orbits will undergo oscillations about the stable orbit if perturbed. To first order, the<br />

restor<strong>in</strong>g force on a bound reduced mass is given by<br />

µ 2 <br />

<br />

= −<br />

2 ( − 0 )=¨ (11.132)<br />

= 0<br />

To the extent that this l<strong>in</strong>ear restor<strong>in</strong>g force dom<strong>in</strong>ates over higher-order terms, then a perturbation of the<br />

stable orbit will undergo simple harmonic oscillations about the stable orbit with angular frequency<br />

v<br />

³ ´<br />

u 2 <br />

t 2 =<br />

=<br />

0<br />

(11.133)<br />

<br />

The above discussion shows that a small amplitude radial oscillation about the stable orbit with amplitude<br />

will be of the form<br />

= s<strong>in</strong>(2 + )<br />

The orbit will be closed if the product of the oscillation frequency and the orbit period is an <strong>in</strong>teger<br />

value.<br />

The fact that planetary orbits <strong>in</strong> the gravitational field are observed to be closed is strong evidence<br />

that the gravitational force field must obey the <strong>in</strong>verse square law. Actually there are small precessions of<br />

planetary orbits due to perturbations of the gravitational field by bodies other than the sun, and due to<br />

relativistic effects. Also the gravitational field near the earth departs slightly from the <strong>in</strong>verse square law<br />

because the earth is not a perfect sphere, and the field does not have perfect spherical symmetry. The study<br />

of the precession of satellites around the earth has been used to determ<strong>in</strong>e the oblate quadrupole and slight<br />

octupole (pear shape) distortion of the shape of the earth.<br />

The most famous test of the <strong>in</strong>verse square law for gravitation is the precession of the perihelion of<br />

Mercury. If the attractive force experienced by Mercury is of the form<br />

F() =− <br />

2+ ˆr<br />

where || is small, then it can be shown that, for approximate circular orbitals, the perihelion will advance<br />

by a small angle per orbit period. That is, the precession is zero if =0, correspond<strong>in</strong>g to an <strong>in</strong>verse<br />

square law dependence which agrees with Bertrand’s theorem. The position of the perihelion of Mercury has<br />

been measured with great accuracy show<strong>in</strong>g that, after correct<strong>in</strong>g for all known perturbations, the perihelion<br />

advances by 43(±5) seconds of arc per century, that is 5 × 10 −7 radians per revolution. This corresponds to<br />

=16 × 10 −7 which is small but still significant. This precession rema<strong>in</strong>ed a puzzle for many years until<br />

1915 when E<strong>in</strong>ste<strong>in</strong> predicted that one consequence of his general theory of relativity is that the planetary<br />

orbit of Mercury should precess at 43 seconds of arc per century, which is <strong>in</strong> remarkable agreement with<br />

observations.


11.10. CLOSED-ORBIT STABILITY 265<br />

11.3 Example: L<strong>in</strong>ear two-body restor<strong>in</strong>g force<br />

The effective potential for a l<strong>in</strong>ear two-body restor<strong>in</strong>g force = − is<br />

At the m<strong>in</strong>imum<br />

Thus<br />

µ<br />

<br />

= 1 2 2 +<br />

<br />

2<br />

2 2<br />

= 0<br />

= − 2<br />

3 =0<br />

µ <br />

2<br />

0 =<br />

<br />

and<br />

µ 2 <br />

<br />

2 = 32<br />

= 0<br />

0<br />

4 + =40<br />

which is a stable orbit. Small perturbations of such a stable circular orbit will have an angular frequency<br />

v<br />

³ ´<br />

u 2 <br />

t 2<br />

=<br />

<br />

1<br />

4<br />

= 0<br />

s<br />

<br />

=2<br />

<br />

Note that this is twice the frequency for the planar harmonic oscillator with the same restor<strong>in</strong>g coefficient.<br />

This is due to the central repulsion, the effective potential well for this rotat<strong>in</strong>g oscillator example has about<br />

half the width for the correspond<strong>in</strong>g planar harmonic oscillator. Note that the k<strong>in</strong>etic energy for the rotational<br />

<br />

motion, which is 2<br />

2<br />

equals the potential energy 1 2 2 2 at the m<strong>in</strong>imum as predicted by the Virial Theorem<br />

for a l<strong>in</strong>ear two-body restor<strong>in</strong>g force.<br />

11.4 Example: Inverse square law attractive force<br />

The effective potential for an <strong>in</strong>verse square law restor<strong>in</strong>g force = − 2 ˆ where isassumedtobe<br />

positive,<br />

At the m<strong>in</strong>imum<br />

Thus<br />

and<br />

µ<br />

<br />

<br />

µ 2 <br />

<br />

2<br />

= − +<br />

2<br />

2 2<br />

= 0<br />

= 2 − 2<br />

3 =0<br />

0 = 2<br />

<br />

= 0<br />

3 0<br />

which is a stable orbit. Small perturbations about such a stable circular orbit will have an angular frequency<br />

v<br />

³ ´<br />

u 2 <br />

t 2 =<br />

=<br />

0<br />

= 2<br />

3<br />

= 0<br />

= 32<br />

4 0<br />

<br />

The k<strong>in</strong>etic energy for oscillations about this stable circular orbit, which is 2<br />

2<br />

equals half the magnitude<br />

2<br />

of the potential energy − <br />

at the m<strong>in</strong>imum as predicted by the Virial Theorem.<br />

− 2<br />

3 0


266 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

11.5 Example: Attractive <strong>in</strong>verse cubic central force<br />

The <strong>in</strong>verse cubic force is an <strong>in</strong>terest<strong>in</strong>g example to <strong>in</strong>vestigate the stability of the orbit equations. One<br />

solution of the <strong>in</strong>verse cubic central force, for a reduced mass is a spiral orbit<br />

= 0 <br />

That this is true can be shown by <strong>in</strong>sert<strong>in</strong>g this orbit <strong>in</strong>to the differential orbit equation.<br />

Us<strong>in</strong>g a B<strong>in</strong>et transformation of the variable to gives<br />

= 1 = 1 0<br />

−<br />

<br />

= − −<br />

0<br />

2 <br />

2 = 2<br />

−<br />

0<br />

Substitut<strong>in</strong>g these <strong>in</strong>to the differential equation of the orbit<br />

gives<br />

That is<br />

2 <br />

2 + = − 2 1<br />

2 ( 1 )<br />

2<br />

− + 1 − = − µ 1<br />

0 0 2 2 0 2 <br />

<br />

µ ¡<br />

1 <br />

= −<br />

<br />

2 +1 ¢ ¡<br />

2<br />

<br />

0 −3<br />

2 +1 ¢ 2<br />

<br />

−3 = −<br />

3<br />

which is a central attractive <strong>in</strong>verse cubic force.<br />

The time dependence of the spiral orbit can be derived s<strong>in</strong>ce the angular momentum gives<br />

˙ =<br />

<br />

2 =<br />

<br />

2 0 2<br />

This can be written as<br />

Integrat<strong>in</strong>g gives<br />

where is a constant. But the orbit gives<br />

2 =<br />

<br />

2 0<br />

<br />

2<br />

2 = <br />

0<br />

2 + <br />

2 = 0 2 2 = 2<br />

<br />

+2<br />

Thus the radius <strong>in</strong>creases or decreases as the square root of the time. That is, an attractive cubic central force<br />

does not have a stable orbit which is what is expected s<strong>in</strong>ce there is no m<strong>in</strong>imum <strong>in</strong> the effective potential<br />

energy. Note that it is obvious that there will be no m<strong>in</strong>imum or maximum for the summation of effective<br />

potential energy s<strong>in</strong>ce, if the force is = − <br />

then the effective potential energy is<br />

3<br />

which has no stable m<strong>in</strong>imum or maximum.<br />

= − <br />

µ <br />

2 2 +<br />

2 <br />

2 1<br />

2 2 = − 2 2


11.10. CLOSED-ORBIT STABILITY 267<br />

11.6 Example: Spirall<strong>in</strong>g mass attached by a str<strong>in</strong>g to a hang<strong>in</strong>g mass<br />

An example of an application of orbit stability is the case shown <strong>in</strong> the adjacent figure. A particle of<br />

mass moves on a horizontal frictionless table. This mass is attached by a light str<strong>in</strong>g of fixed length and<br />

rotates about a hole <strong>in</strong> the table. The str<strong>in</strong>g is attached to a second equal mass that is hang<strong>in</strong>g vertically<br />

downwards with no angular motion.<br />

The equations are most conveniently expressed <strong>in</strong> cyl<strong>in</strong>drical<br />

coord<strong>in</strong>ates ( ) with the orig<strong>in</strong> at the hole <strong>in</strong> the table, and <br />

vertically upward. The fixedlengthofthestr<strong>in</strong>grequires = −.<br />

z<br />

The potential energy is<br />

= = ( − )<br />

O<br />

The system is central and conservative, thus the Hamiltonian<br />

can be written as<br />

= ³<br />

˙ 2 + ˙2´<br />

2 + <br />

2 + ( − ) =<br />

2<br />

2<br />

The Lagrangian is <strong>in</strong>dependent of , thatis, is cyclic, thus the<br />

angular momentum 2 ˙ = is a constant of motion. Substitut<strong>in</strong>g<br />

this <strong>in</strong>to the Hamiltonian equation gives<br />

Rotat<strong>in</strong>g mass on a frictionless<br />

horizontal table connected to a<br />

suspended mass .<br />

˙ 2 +<br />

2<br />

+ ( − ) =<br />

22 The effective potential is<br />

=<br />

2<br />

+ ( − )<br />

22 which is shown <strong>in</strong> the adjacent figure. The stationary value occurs when<br />

µ<br />

<br />

<br />

0<br />

= − 2<br />

3 0<br />

+ =0<br />

That is, when the angular momentum is related to the radius by<br />

2 = 2 3 0<br />

Note that 0 =0if =0.<br />

The stability of the solution is given by the second derivative<br />

µ 2 <br />

2 <br />

0<br />

= 32<br />

4 0<br />

= 3<br />

0<br />

0<br />

Therefore the stationary po<strong>in</strong>t is stable.<br />

Note that the equation of motion for the m<strong>in</strong>imum can be<br />

expressed <strong>in</strong> terms of the restor<strong>in</strong>g force on the two masses<br />

µ 2 <br />

<br />

2¨ = −<br />

2 ( − 0 )<br />

0<br />

Thus the system undergoes harmonic oscillation with frequency<br />

s<br />

3 r<br />

<br />

= 0 3<br />

2 = 2 0<br />

The solution of this system is stable and undergoes simple<br />

harmonic motion.<br />

Effective potential for two connected masses.<br />

r


268 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

11.11 The three-body problem<br />

Two bodies <strong>in</strong>teract<strong>in</strong>g via conservative central forces can be<br />

solved analytically for the <strong>in</strong>verse square law and the Hooke’s law<br />

radial dependences as already discussed. Central forces that have<br />

other radial dependences for the equations of motion may not be<br />

expressible <strong>in</strong> terms of simple functions, nevertheless the motion<br />

always can be given <strong>in</strong> terms of an <strong>in</strong>tegral. For a gravitational<br />

system compris<strong>in</strong>g ≥ 3 bodies that are <strong>in</strong>teract<strong>in</strong>g via the twobody<br />

central gravitational force, then the equations of motion<br />

can be written as<br />

¨q = G<br />

X<br />

<br />

6=<br />

<br />

(q − q )<br />

|q − q | 3 ( =1 2 )<br />

Even when all the bodies are <strong>in</strong>teract<strong>in</strong>g via two-body central<br />

forces, the problem usually is <strong>in</strong>soluble <strong>in</strong> terms of known analytic<br />

<strong>in</strong>tegrals. Newton first posed the difficulty of the three-body<br />

Kepler problem which has been studied extensively by mathematicians<br />

and physicists. No known general analytic <strong>in</strong>tegral<br />

solution has been found. Each body for the -body system has<br />

6 degrees of freedom, that is, 3 for position and 3 for momentum.<br />

The center-of-mass motion can be factored out, therefore<br />

the center-of-mass system for the -body system has 6 − 10 degrees<br />

of freedom after subtraction of 3 degrees for location of the<br />

center of mass, 3 for the l<strong>in</strong>ear momentum of the center of mass,<br />

3 for rotation of the center of mass, and 1 for the total energy of<br />

Figure 11.10: A contour plot of the effective<br />

potential for the Sun-Earth gravitational<br />

system <strong>in</strong> the rotat<strong>in</strong>g frame where<br />

the Sun and Earth are stationary. The<br />

5 Lagrange po<strong>in</strong>ts are saddle po<strong>in</strong>ts<br />

where the net force is zero. (Figure created<br />

by NASA)<br />

the system. Thus for =2there are 12 − 10 = 2 degrees of freedom for the two-body system for which the<br />

Kepler approach takes to be r and For =3there are 8 degrees of freedom <strong>in</strong> the center of mass system<br />

that have to be determ<strong>in</strong>ed.<br />

Numerical solutions to the three-body problem can be obta<strong>in</strong>ed us<strong>in</strong>g successive approximation or perturbation<br />

methods <strong>in</strong> computer calculations. The problem can be simplified by restrict<strong>in</strong>g the motion to<br />

either of follow<strong>in</strong>g two approximations:<br />

1) Planar approximation<br />

This approximation assumes that the three masses move <strong>in</strong> the same plane, that is, the number of degrees<br />

of freedom are reduced from 8 to 6 which simplifies the numerical solution.<br />

2) Restricted three-body approximation<br />

The restricted three-body approximation assumes that two of the masses are large and bound while the<br />

third mass is negligible such that the perturbation of the motion of the larger two by the third body is<br />

negligible. This approximation essentially reduces the system to a two body problem <strong>in</strong> order to calculate<br />

the gravitational fields that act on the third much lighter mass.<br />

Euler and Lagrange showed that the restricted three-body system has five po<strong>in</strong>ts at which the comb<strong>in</strong>ed<br />

gravitational attraction plus centripetal force of the two large bodies cancel. These are called the Lagrange<br />

po<strong>in</strong>ts and are used for park<strong>in</strong>g satellites <strong>in</strong> stable orbits with respect to the Earth-Moon system, or with<br />

respect to the Sun-Earth system. Figure 1110 illustrates the five Lagrange po<strong>in</strong>ts for the Earth-Sun system.<br />

Only two of the Lagrange po<strong>in</strong>ts, 4 and 5 lead to stable orbits. Note that these Lagrange po<strong>in</strong>ts are fixed<br />

with respect to the Earth-Sun system which rotates with respect to <strong>in</strong>ertial coord<strong>in</strong>ate frames. The 1900’s<br />

discovery of the Trojan asteroids at the 4 and 5 Lagrange po<strong>in</strong>ts of the Sun-Jupiter system confirmed the<br />

Lagrange predictions.<br />

Po<strong>in</strong>caré showed that the motion of a light mass bound to two heavy bodies can exhibit extreme sensitivity<br />

to <strong>in</strong>itial conditions as well as characteristics of chaos. Solution of the three-body problem has rema<strong>in</strong>ed a<br />

largely unsolved problem s<strong>in</strong>ce Newton identified the difficulties <strong>in</strong>volved.


11.12. TWO-BODY SCATTERING 269<br />

11.12 Two-body scatter<strong>in</strong>g<br />

Two mov<strong>in</strong>g bodies, that are <strong>in</strong>teract<strong>in</strong>g via a central force, scatter when the force is repulsive, or when<br />

an attractive system is unbound. Two-body scatter<strong>in</strong>g of bodies is encountered extensively <strong>in</strong> the fields of<br />

astronomy, atomic, nuclear, and particle physics. The probability of such scatter<strong>in</strong>g is most conveniently<br />

expressed <strong>in</strong> terms of scatter<strong>in</strong>g cross sections def<strong>in</strong>ed below.<br />

11.12.1 Total two-body scatter<strong>in</strong>g cross section<br />

The concept of scatter<strong>in</strong>g cross section for two-body scatter<strong>in</strong>g<br />

is most easily described for the total two-body cross<br />

section. The probability that a beam of <strong>in</strong>cident po<strong>in</strong>t<br />

particles/second, distributed over a cross sectional area <br />

will hit a s<strong>in</strong>gle solid object, hav<strong>in</strong>g a cross sectional area <br />

is given by the ratio of the areas as illustrated <strong>in</strong> figure 1111.<br />

That is,<br />

=<br />

<br />

(11.134)<br />

<br />

where it is assumed that For a spherical target<br />

body of radius , the cross section = 2 The scatter<strong>in</strong>g<br />

probability is proportional to the cross section which<br />

is the cross section of the target body perpendicular to the<br />

beam; thus has the units of area.<br />

A B<br />

Figure 11.11: Scatter<strong>in</strong>g probability for an<br />

<strong>in</strong>cident beam of cross sectional area A by a<br />

target body of cross sectional area .<br />

S<strong>in</strong>ce the <strong>in</strong>cident beam of <strong>in</strong>cident po<strong>in</strong>t particles/second,<br />

has a cross sectional area ,thenitwillhave<br />

an areal density given by<br />

= <br />

beam particles 2 / sec (11.135)<br />

<br />

The number of beam particles scattered per second by this s<strong>in</strong>gle target scatterer equals<br />

= =<br />

= (11.136)<br />

<br />

Thus the cross section for scatter<strong>in</strong>g by this s<strong>in</strong>gle target body is<br />

= <br />

<br />

Scattered particles/sec<br />

=<br />

<strong>in</strong>cident beam/m 2 /sec<br />

Realistically one will have many target scatterers <strong>in</strong> the target and the total scatter<strong>in</strong>g probability <strong>in</strong>creases<br />

proportionally to the number of target scatterers. That is, for a target compris<strong>in</strong>g an areal density of <br />

target bodies per unit area of the <strong>in</strong>cident beam, then the number scattered will <strong>in</strong>crease proportional to the<br />

target areal density That is, there will be scatter<strong>in</strong>g bodies that <strong>in</strong>teract with the beam assum<strong>in</strong>g<br />

that the target has a larger area than the beam. Thus the total number scattered per second by a target<br />

that comprises multiple scatterers is<br />

= <br />

<br />

= (11.137)<br />

<br />

Note that this is <strong>in</strong>dependent of the cross sectional area of the beam assum<strong>in</strong>g that the target area is larger<br />

than that of the beam. That is, the number scattered per second is proportional to the cross section times<br />

the product of the number of <strong>in</strong>cident particles per second, and the areal density of target scatterers,<br />

. Typical cross sections encountered <strong>in</strong> astrophysics are ≈ 10 14 2 ,<strong>in</strong>atomicphysics: ≈ 10 −20 2 ,<br />

and <strong>in</strong> nuclear physics; ≈ 10 −28 2 = 3<br />

N. B., the above proof assumed that the target size is larger than the cross sectional area of the <strong>in</strong>cident<br />

beam. If the size of the target is smaller than the beam, then is replaced by the areal density/sec of the<br />

beam and is replaced by the number of target particles and the cross-sectional size of the target<br />

cancels.<br />

3 The term "barn" was chosen because nuclear physicists joked that the cross sections for neutron scatter<strong>in</strong>g by nuclei were<br />

as large as a barn door.


270 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

11.12.2 Differential two-body scatter<strong>in</strong>g cross section<br />

The differential two-body scatter<strong>in</strong>g cross section gives much<br />

more detailed <strong>in</strong>formation of the scatter<strong>in</strong>g force than does<br />

the total cross section because of the correlation between the<br />

impact parameter and the scatter<strong>in</strong>g angle. That is, a measurement<br />

of the number of beam particles scattered <strong>in</strong>to a<br />

given solid angle as a function of scatter<strong>in</strong>g angles probes<br />

the radial form of the scatter<strong>in</strong>g force.<br />

The differential cross section for scatter<strong>in</strong>g of an <strong>in</strong>cident<br />

beam by a s<strong>in</strong>gle target body <strong>in</strong>to a solid angle Ω at scatter<strong>in</strong>g<br />

angles is def<strong>in</strong>ed to be<br />

<br />

Ω () ≡ 1 ( )<br />

(11.138)<br />

Ω<br />

where the right-hand side is the ratio of the number scattered<br />

per target nucleus <strong>in</strong>to solid angle Ω( ) to the <strong>in</strong>cident<br />

beam <strong>in</strong>tensity 2 .<br />

Similar reason<strong>in</strong>g used to derive equation 11137 leads to<br />

the number of beam particles scattered <strong>in</strong>to a solid angle<br />

Ω for beam particles <strong>in</strong>cident upon a target with areal<br />

density is<br />

( )<br />

Ω<br />

db<br />

b<br />

Figure 11.12: The equivalent one-body problem<br />

for scatter<strong>in</strong>g of a reduced mass by a<br />

force centre <strong>in</strong> the centre of mass system.<br />

<br />

= () (11.139)<br />

Ω<br />

Consider the equivalent one-body system for scatter<strong>in</strong>g of one body by a scatter<strong>in</strong>g force center <strong>in</strong><br />

the center of mass. As shown <strong>in</strong> figures 116 and 1112, the perpendicular distance between the center of<br />

force of the two body system and trajectory of the <strong>in</strong>com<strong>in</strong>g body at <strong>in</strong>f<strong>in</strong>ite distance is called the impact<br />

parameter . For a central force the scatter<strong>in</strong>g system has cyl<strong>in</strong>drical symmetry, therefore the solid angle<br />

Ω() =s<strong>in</strong> can be <strong>in</strong>tegrated over the azimuthal angle to give Ω() =2 s<strong>in</strong> <br />

For the <strong>in</strong>verse-square, two-body, central force there is a one-to-one correspondence between impact<br />

parameter and scatter<strong>in</strong>g angle for a given bombard<strong>in</strong>g energy. In this case, assum<strong>in</strong>g conservation of<br />

flux means that the <strong>in</strong>cident beam particles pass<strong>in</strong>g through the impact-parameter annulus between and<br />

+ must equal the the number pass<strong>in</strong>g between the correspond<strong>in</strong>g angles and + That is, for an<br />

<strong>in</strong>cident beam flux of 2 the number of particles per second pass<strong>in</strong>g through the annulus is<br />

2 || =2 s<strong>in</strong> || (11.140)<br />

Ω<br />

The modulus is used to ensure that the number of particles is always positive. Thus<br />

<br />

Ω = <br />

s<strong>in</strong> <br />

11.12.3 Impact parameter dependence on scatter<strong>in</strong>g angle<br />

<br />

¯<br />

¯ (11.141)<br />

If the function = ( ) is known, then it is possible to evaluate ¯¯ ¯<br />

<br />

which can be used <strong>in</strong> equation<br />

11141 to calculate the differential cross section. A simple and important case to consider is two-body elastic<br />

scatter<strong>in</strong>g for the <strong>in</strong>verse-square law force such as the Coulomb or gravitational forces. To avoid confusion <strong>in</strong><br />

the follow<strong>in</strong>g discussion, the center-of-mass scatter<strong>in</strong>g angle will be called while the angle used to def<strong>in</strong>e<br />

the hyperbolic orbits <strong>in</strong> the discussion of trajectories for the <strong>in</strong>verse square law, will be called .<br />

In chapter 118 the equivalent one-body representation gave that the radial distance for a trajectory for<br />

the <strong>in</strong>verse square law is given by<br />

1<br />

= − [1 + cos ] (11.142)<br />

2 Note that closest approach occurs when =0while for →∞the bracket must equal zero, that is<br />

cos ∞ = ±<br />

1<br />

¯ ¯ (11.143)


11.12. TWO-BODY SCATTERING 271<br />

The polar angle is measured with respect to the symmetry axis of the two-body system which is along<br />

the l<strong>in</strong>e of distance of closest approach as shown <strong>in</strong> figure 116. The geometry and symmetry show that the<br />

scatter<strong>in</strong>g angle is related to the trajectory angle ∞ by<br />

= − 2 ∞ (11.144)<br />

Equation 1150 gives that<br />

∞ =<br />

Z ∞<br />

m<strong>in</strong><br />

±<br />

³<br />

´ (11.145)<br />

<br />

r2<br />

2 − −<br />

2<br />

2 2<br />

S<strong>in</strong>ce<br />

2 = 2 2 = 2 2 (11.146)<br />

then the scatter<strong>in</strong>g angle can be written as.<br />

∞ = − Z ∞<br />

<br />

= r<br />

2<br />

³ ´ (11.147)<br />

m<strong>in</strong><br />

2 1 −<br />

<br />

<br />

− 2<br />

2<br />

Let = 1 ,then<br />

∞ = − <br />

2<br />

=<br />

Z ∞<br />

m<strong>in</strong><br />

<br />

r ³ 2´ (11.148)<br />

1 −<br />

<br />

<br />

− 2 <br />

For the repulsive <strong>in</strong>verse square law<br />

= − <br />

= − (11.149)<br />

where is taken to be positive for a repulsive force. Thus the scatter<strong>in</strong>g angle relation becomes<br />

∞ = − <br />

2<br />

=<br />

Z ∞<br />

m<strong>in</strong><br />

<br />

r ³ 2´ (11.150)<br />

1+ <br />

<br />

− 2 <br />

The solution of this equation is given by equation 1163 to be<br />

= 1 = − [1 + cos ] (11.151)<br />

2 where the eccentricity<br />

=<br />

s<br />

1+ 2 2<br />

2 (11.152)<br />

For →∞=0then, as shown previously,<br />

1<br />

¯ ¯ =cos ∞ =cos − =s<strong>in</strong> 2 2<br />

(11.153)<br />

Therefore<br />

2 <br />

= p <br />

<br />

2 − 1=cot (11.154)<br />

2<br />

that is, the impact parameter is given by the relation<br />

=<br />

<br />

2 <br />

cot 2<br />

(11.155)<br />

Thus, for an <strong>in</strong>verse-square law force, the two-body scatter<strong>in</strong>g<br />

has a one-to-one correspondence between impact parameter <br />

and scatter<strong>in</strong>g angle as shown schematically <strong>in</strong> figure 1113.<br />

Figure 11.13: Impact parameter dependence<br />

on scatter<strong>in</strong>g angle for Rutherford<br />

scatter<strong>in</strong>g.


272 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

If is negative, which corresponds to an attractive <strong>in</strong>verse square<br />

law, then one gets the same relation between impact parameter and<br />

scatter<strong>in</strong>g angle except that the sign of the impact parameter is<br />

opposite. This means that the hyperbolic trajectory has an <strong>in</strong>terior<br />

rather than exterior focus. That is, the trajectory partially orbits<br />

around the center of force rather than be<strong>in</strong>g repelled away.<br />

Note that the distance of closest approach is related to the<br />

eccentricity by equation 11151, therefore<br />

m<strong>in</strong> =<br />

Ã<br />

m<strong>in</strong> =<br />

1+ 1<br />

2 s<strong>in</strong> 2<br />

Note that for = 180 then<br />

=<br />

<br />

2 <br />

(1 + ) (11.156)<br />

!<br />

(11.157)<br />

<br />

m<strong>in</strong><br />

= ( m<strong>in</strong>) (11.158)<br />

Figure 11.14: <strong>Classical</strong> trajectories<br />

for scatter<strong>in</strong>g to a given angle by the<br />

repulsive Coulomb field plus the attractive<br />

nuclear field for three different<br />

impact parameters. Path 1 is<br />

pure Coulomb. Paths 2 and 3 <strong>in</strong>clude<br />

Coulomb plus nuclear <strong>in</strong>teractions.<br />

The dashed parts of trajectories2and3correspondtoonlythe<br />

Coulomb force act<strong>in</strong>g, i.e. zero nuclear<br />

force<br />

which is what you would expect from equat<strong>in</strong>g the <strong>in</strong>cident k<strong>in</strong>etic<br />

energy to the potential energy at the distance of closest approach.<br />

For scatter<strong>in</strong>g of two nuclei by the repulsive Coulomb force, if the<br />

impact parameter becomes small enough, the attractive nuclear force<br />

also acts lead<strong>in</strong>g to impact-parameter dependent effective potentials<br />

illustrated <strong>in</strong> figure 1114 Trajectory 1 does not overlap the nuclear<br />

force and thus is pure Coulomb. Trajectory 2 <strong>in</strong>teracts at the periphery<br />

of the nuclear potential and the trajectory deviates from pure Coulomb shown dashed. Trajectory 3<br />

passes through the <strong>in</strong>terior of the nuclear potential. These three trajectories all can lead to the same scatter<strong>in</strong>g<br />

angle and thus there no longer is a one-to-one correspondence between scatter<strong>in</strong>g angle and impact<br />

parameter.<br />

11.12.4 Rutherford scatter<strong>in</strong>g<br />

Two models of the nucleus evolved <strong>in</strong> the 1900’s, the Rutherford model assumed electrons orbit<strong>in</strong>g around a<br />

small nucleus like planets around the sun, while J.J. Thomson’s ”plum-pudd<strong>in</strong>g” model assumed the electrons<br />

were embedded <strong>in</strong> a uniform sphere of positive charge the size of the atom. When Rutherford derived his<br />

classical formula <strong>in</strong> 1911 he realized that it can be used to determ<strong>in</strong>e the size of the nucleus s<strong>in</strong>ce the electric<br />

field obeys the <strong>in</strong>verse square law only when outside of the charged spherical nucleus. Inside a uniform sphere<br />

of charge the electric field is E ∝ r and thus the scatter<strong>in</strong>g cross section will not obey the Rutherford relation<br />

for distances of closest approach that are less than the radius of the sphere of negative charge. Observation<br />

of the angle beyond which the Rutherford formula breaks down immediately determ<strong>in</strong>es the radius of the<br />

nucleus.<br />

<br />

For pure Coulomb scatter<strong>in</strong>g, equation 11155 canbeusedtoevaluate¯¯ ¯<br />

<br />

whichwhenused<strong>in</strong>equation<br />

11141 gives the center-of-mass Rutherford scatter<strong>in</strong>g cross section<br />

<br />

Ω = 1 µ 2 1<br />

4 2 s<strong>in</strong> 4 2<br />

(11.159)<br />

This cross section assumes elastic scatter<strong>in</strong>g by a repulsive two-body <strong>in</strong>verse-square central force. For scatter<strong>in</strong>g<br />

of nuclei <strong>in</strong> the Coulomb potential, the constant is given to be<br />

= 2<br />

4 <br />

(11.160)<br />

The cross section, scatter<strong>in</strong>g angle and of equation 11159 are evaluated <strong>in</strong> the center-of-mass coord<strong>in</strong>ate<br />

system, whereas usually two-body elastic scatter<strong>in</strong>g data <strong>in</strong>volve scatter<strong>in</strong>g of the projectiles by a<br />

stationary target as discussed <strong>in</strong> chapter 1113


11.12. TWO-BODY SCATTERING 273<br />

Gieger and Marsden performed scatter<strong>in</strong>g of 77 MeV particles from a th<strong>in</strong> gold foil and proved that<br />

the differential scatter<strong>in</strong>g cross section obeyed the Rutherford formula back to angles correspond<strong>in</strong>g to a<br />

distance of closest approach of 10 −14 which is much smaller that the 10 −10 size of the atom. This<br />

validated the Rutherford model of the atom and immediately led to the Bohr model of the atom which<br />

played such a crucial role <strong>in</strong> the development of quantum mechanics. Bohr showed that the agreement with<br />

the Rutherford formula implies the Coulomb field obeys the <strong>in</strong>verse square law to small distances. This work<br />

was performed at Manchester University, England between 1908 and 1913. It is fortunate that the classical<br />

result is identical to the quantal cross section for scatter<strong>in</strong>g, otherwise the development of modern physics<br />

could have been delayed for many years.<br />

Scatter<strong>in</strong>g of very heavy ions, such as 208 Pb, can electromagnetically excite target nuclei. For the Coulomb<br />

force the impact parameter and the distance of closest approach, m<strong>in</strong> are directly related to the scatter<strong>in</strong>g<br />

angle by equation 11155. Thus observ<strong>in</strong>g the angle of the scattered projectile unambiguously determ<strong>in</strong>es<br />

the hyperbolic trajectory and thus the electromagnetic impulse given to the collid<strong>in</strong>g nuclei. This process,<br />

called Coulomb excitation, uses the measured angular distribution of the scattered ions for <strong>in</strong>elastic excitation<br />

of the nuclei to precisely and unambiguously determ<strong>in</strong>e the Coulomb excitation cross section as a function<br />

of impact parameter. This unambiguously determ<strong>in</strong>es the shape of the nuclear charge distribution.<br />

11.7 Example: Two-body scatter<strong>in</strong>g by an <strong>in</strong>verse cubic force<br />

Assume two-body scatter<strong>in</strong>g by a potential = <br />

where 0. This corresponds to a repulsive two-body<br />

2<br />

force F = 2<br />

<br />

ˆr. Insert this force <strong>in</strong>to B<strong>in</strong>et’s differential orbit, equation 1139 gives<br />

3<br />

2 µ<br />

<br />

2 + 1+ 2 <br />

2 =0<br />

The solution is of the form = s<strong>in</strong>( + ) where and are constants of <strong>in</strong>tegration, = 2 ˙ and<br />

µ<br />

2 = 1+ 2 <br />

2<br />

Initially = ∞, =0 and therefore =0. Also at = ∞, = 1 2 ˙2 ∞ ,thatis| ˙ ∞ | =<br />

q<br />

2<br />

. Then<br />

˙ = <br />

˙ = <br />

2 = − <br />

= − cos ()<br />

<br />

The <strong>in</strong>itial energy gives that = 1 √<br />

2 Hence the orbit equation is<br />

= 1 = √ 2<br />

<br />

s<strong>in</strong> ()<br />

The above trajectory has a distance of closest approach, m<strong>in</strong> , when m<strong>in</strong> = <br />

symmetry of the orbit, the scatter<strong>in</strong>g angle is given by<br />

µ<br />

= − 2 0 = 1 − 1 <br />

<br />

2 .<br />

Moreover, due to the<br />

S<strong>in</strong>ce 2 = 2 2 ˙ ∞ 2 =2 2 then<br />

1 − µ<br />

= 1+ 2 − 1<br />

2<br />

=<br />

µ1+ − 1<br />

2<br />

2 2 <br />

This gives that the impact parameter is related to scatter<strong>in</strong>g angle by<br />

2 = ( − ) 2<br />

(2 − ) <br />

This impact parameter relation can be used <strong>in</strong> equation 11141 to give the differential cross section<br />

<br />

Ω = <br />

<br />

s<strong>in</strong> ¯<br />

¯ = 2 ( − )<br />

(2 − ) 2 2<br />

These orbits are called Cotes spirals.


274 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

11.13 Two-body k<strong>in</strong>ematics<br />

So far the discussion has been restricted to the center-of-momentum system. Actual scatter<strong>in</strong>g measurements<br />

are performed <strong>in</strong> the laboratory frame, and thus it is necessary to transform the scatter<strong>in</strong>g angle, energies<br />

and cross sections between the laboratory and center-of-momentum coord<strong>in</strong>ate frame. In pr<strong>in</strong>ciple the<br />

transformation between the center-of-momentum and laboratory frames is straightforward, us<strong>in</strong>g the vector<br />

addition of the center-of-mass velocity vector and the center-of-momentum velocity vectors of the two bodies.<br />

The follow<strong>in</strong>g discussion assumes non-relativistic k<strong>in</strong>ematics apply.<br />

In chapter 28 it was shown that, for Newtonian mechanics, the center-of-mass and center-of-momentum<br />

frames of reference are identical. By def<strong>in</strong>ition, <strong>in</strong> the center-of-momentum frame the vector sum of the<br />

l<strong>in</strong>ear momentum of the <strong>in</strong>com<strong>in</strong>g projectile <br />

and target, <br />

are equal and opposite. That is<br />

p <br />

<br />

+ p <br />

=0 (11.161)<br />

Us<strong>in</strong>g the center-of-momentum frame, coupled with the conservation of l<strong>in</strong>ear momentum, implies that the<br />

vector sum of the f<strong>in</strong>al momenta of the reaction products, <br />

also is zero. That is<br />

X<br />

=1<br />

p <br />

=0 (11.162)<br />

An additional constra<strong>in</strong>t is that energy conservation relates the <strong>in</strong>itial and f<strong>in</strong>al k<strong>in</strong>etic energies by<br />

¡ ¢<br />

<br />

2<br />

<br />

+<br />

2 <br />

¡ ¢<br />

<br />

2<br />

<br />

+ =<br />

2 <br />

¡ ¢<br />

<br />

2<br />

<br />

+<br />

2 <br />

¡ ¢<br />

<br />

2<br />

<br />

(11.163)<br />

2 <br />

where the value is the energy contributed to the f<strong>in</strong>al total k<strong>in</strong>etic energy by the reaction between the<br />

<strong>in</strong>com<strong>in</strong>g projectile and target. For exothermic reactions, 0 the summed k<strong>in</strong>etic of the reaction products<br />

exceeds the sum of the <strong>in</strong>com<strong>in</strong>g k<strong>in</strong>etic energies, while for endothermic reactions, 0 the summed k<strong>in</strong>etic<br />

energy of the reaction products is less than that of the <strong>in</strong>com<strong>in</strong>g channel.<br />

For two-body k<strong>in</strong>ematics, the follow<strong>in</strong>g are three advantages to work<strong>in</strong>g <strong>in</strong> the center-of-momentum frame<br />

of reference.<br />

1. The two <strong>in</strong>cident collid<strong>in</strong>g bodies are col<strong>in</strong>ear as are the two f<strong>in</strong>al bodies.<br />

2. The l<strong>in</strong>ear momenta for the two collid<strong>in</strong>g bodies are identical <strong>in</strong> both the <strong>in</strong>cident channel and the<br />

outgo<strong>in</strong>g channel.<br />

3. The total energy <strong>in</strong> the center-of-momentum coord<strong>in</strong>ate frame is the energy available to the reaction<br />

dur<strong>in</strong>g the collision. The trivial k<strong>in</strong>etic energy of the center-of-momentum frame relative to the<br />

laboratory frame is handled separately.<br />

The k<strong>in</strong>ematics for two-body reactions is easily determ<strong>in</strong>ed us<strong>in</strong>g the conservation of l<strong>in</strong>ear momentum<br />

along and perpendicular to the beam direction plus the conservation of energy, 11161 − 11163. Notethatit<br />

is common practice to use the term “center-of-mass” rather than “center-of-momentum” <strong>in</strong> spite of the fact<br />

that, for relativistic mechanics, only the center-of-momentum is a mean<strong>in</strong>gful concept.<br />

General features of the transformation between the center-of-momentum and laboratory frames of reference<br />

are best illustrated by elastic or <strong>in</strong>elastic scatter<strong>in</strong>g of nuclei where the two reaction products <strong>in</strong> the f<strong>in</strong>al<br />

channel are identical to the <strong>in</strong>cident bodies. Inelastic excitation of an excited state energy of ∆ <strong>in</strong> either<br />

reaction product corresponds to = −∆ while elastic scatter<strong>in</strong>g corresponds to = −∆ =0.<br />

For <strong>in</strong>elastic scatter<strong>in</strong>g, the conservation of l<strong>in</strong>ear momenta for the outgo<strong>in</strong>g channel <strong>in</strong> the center-ofmomentum<br />

simplifies to<br />

p <br />

<br />

+ p <br />

=0 (11.164)<br />

that is, the l<strong>in</strong>ear momenta of the two reaction products are equal and opposite.<br />

Assume that the center-of-momentum direction of the scattered projectile is at an angle = relative<br />

to the direction of the <strong>in</strong>com<strong>in</strong>g projectile and that the scattered target nucleus is scattered at a centerof-momentum<br />

direction = − .<br />

magnitudes of the <strong>in</strong>com<strong>in</strong>g and outgo<strong>in</strong>g projectile momenta are equal, that is, ¯¯ <br />

Elastic scatter<strong>in</strong>g corresponds to simple scatter<strong>in</strong>g for which the<br />

¯ ¯¯<br />

¯<br />

= <br />

¯.


11.13. TWO-BODY KINEMATICS 275<br />

Figure 11.15: Vector hodograph of the scattered projectile and target velocities for a projectile, with <strong>in</strong>cident<br />

velocity that is elastically scattered by a stationary target body. The circles show the magnitude of the<br />

projectile and target body f<strong>in</strong>al velocities <strong>in</strong> the center of mass. The center-of-mass velocity vectors are<br />

shown as dashed l<strong>in</strong>es while the laboratory vectors are shown as solid l<strong>in</strong>es. The left hodograph shows<br />

normal k<strong>in</strong>ematics where the projectile mass is less than the target mass. The right hodograph shows<br />

<strong>in</strong>verse k<strong>in</strong>ematics where the projectile mass is greater than the target mass. For elastic scatter<strong>in</strong>g = 0 .<br />

Velocities<br />

The transformation between the center-of-momentum and laboratory frames requires knowledge of the particle<br />

velocities which can be derived from the l<strong>in</strong>ear momenta s<strong>in</strong>ce the particle masses are known. Assume<br />

that a projectile, mass ,with<strong>in</strong>cidentenergy <strong>in</strong> the laboratory frame bombards a stationary target<br />

with mass The <strong>in</strong>cident projectile velocity is given by<br />

=<br />

The <strong>in</strong>itial velocities <strong>in</strong> the laboratory frame are taken to be<br />

The f<strong>in</strong>al velocities <strong>in</strong> the laboratory frame after the <strong>in</strong>elastic collision are<br />

r<br />

2<br />

<br />

(11.165)<br />

= (Initial Lab velocities)<br />

= 0<br />

0 <br />

(F<strong>in</strong>al Lab velocities)<br />

0 <br />

In the center-of-momentum coord<strong>in</strong>ate system, equation 1110 implies that the <strong>in</strong>itial center-of-momentum<br />

velocities are<br />

= <br />

<br />

+ <br />

= <br />

<br />

+ <br />

(11.166)<br />

It is simple to derive that the f<strong>in</strong>al center-of-momentum velocities after the <strong>in</strong>elastic collision are given


276 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

by<br />

0 =<br />

0 =<br />

<br />

+ <br />

r<br />

2<br />

<br />

˜<br />

r<br />

2<br />

˜ (11.167)<br />

+ <br />

The energy ˜ is def<strong>in</strong>ed to be given by<br />

˜ = + (1 + <br />

) (11.168)<br />

<br />

where = −∆ which is the excitation energy of the f<strong>in</strong>al excited states <strong>in</strong> the outgo<strong>in</strong>g channel.<br />

Angles<br />

The angles of the scattered recoils are written as<br />

<br />

(F<strong>in</strong>al laboratory angles)<br />

<br />

and<br />

= (F<strong>in</strong>al CM angles)<br />

= − <br />

where is the center-of-mass (center-of-momentum) scatter<strong>in</strong>g angle.<br />

Figure 1115 shows that the angle relations between the laboratory and center of momentum frames for<br />

the scattered projectile are connected by<br />

s<strong>in</strong>( − )<br />

s<strong>in</strong> <br />

= <br />

<br />

r<br />

˜<br />

≡ (11.169)<br />

where<br />

= 1<br />

q<br />

<br />

1+ <br />

(1 + <br />

<br />

)<br />

= <br />

1<br />

q<br />

1+ <br />

<br />

( + <br />

<br />

)<br />

(11.170)<br />

and <br />

<br />

is the energy per nucleon on the <strong>in</strong>cident projectile.<br />

Equation 11169 can be rewritten as<br />

tan =<br />

s<strong>in</strong> <br />

cos + <br />

(11.171)<br />

Another useful relation from equation 11169 gives the center-of-momentum scatter<strong>in</strong>g angle <strong>in</strong> terms of<br />

the laboratory scatter<strong>in</strong>g angle.<br />

=s<strong>in</strong> −1 ( s<strong>in</strong> )+ (11.172)<br />

This gives the difference <strong>in</strong> angle between the lab scatter<strong>in</strong>g angle and the center-of-momentum scatter<strong>in</strong>g<br />

angle. Be careful with this relation s<strong>in</strong>ce is two-valued for <strong>in</strong>verse k<strong>in</strong>ematics correspond<strong>in</strong>g to the two<br />

possible signs for the solution.<br />

The angle relations between the lab and center-of-momentum for the recoil<strong>in</strong>g target nucleus are connected<br />

by<br />

That is<br />

r<br />

s<strong>in</strong>( − )<br />

s<strong>in</strong> = ≡ ˜ (11.173)<br />

<br />

˜<br />

=s<strong>in</strong> −1 (˜ s<strong>in</strong> )+ (11.174)


11.13. TWO-BODY KINEMATICS 277<br />

Figure 11.16: The k<strong>in</strong>ematic correlation of the laboratory and center-of-mass scatter<strong>in</strong>g angles of the recoil<strong>in</strong>g<br />

projectile and target nuclei for scatter<strong>in</strong>g for 43/nucleon 104 Pd on 208 Pb (left) and for the <strong>in</strong>verse<br />

43/nucleon 208 Pb on 104 Pd (right). The projectile scatter<strong>in</strong>g angles are shown by solid l<strong>in</strong>es while the<br />

recoil<strong>in</strong>g target angles are shown by dashed l<strong>in</strong>es. The blue curves correspond to elastic scatter<strong>in</strong>g, that is<br />

=0 while the red curves correspond to <strong>in</strong>elastic scatter<strong>in</strong>g with = −5.<br />

where<br />

1<br />

˜ = q<br />

=<br />

1+ <br />

(1 + <br />

<br />

)<br />

1<br />

q<br />

1+ <br />

<br />

( + <br />

<br />

)<br />

(11.175)<br />

Note that ˜ is the same under <strong>in</strong>terchange of the two nuclei at the same <strong>in</strong>cident energy/nucleon, and<br />

that ˜ is always larger than or equal to unity s<strong>in</strong>ce is negative. For elastic scatter<strong>in</strong>g ˜ =1which gives<br />

= 1 ( − )<br />

2<br />

(Recoil lab angle for elastic scatter<strong>in</strong>g)<br />

For the target recoil equation 11173 can be rewritten as<br />

tan =<br />

s<strong>in</strong> <br />

cos +˜<br />

(Target lab to CM angle conversion)<br />

Velocity vector hodographs provide useful <strong>in</strong>sight <strong>in</strong>to the behavior of the k<strong>in</strong>ematic solutions. As shown<br />

<strong>in</strong> figure 1115, <strong>in</strong> the center-of-momentum frame the scattered projectile has a fixed f<strong>in</strong>al velocity 0 ,that<br />

is, the velocity vector describes a circle as a function of . The vector addition of this vector and the velocity<br />

of the center-of-mass vector − gives the laboratory frame velocity 0 . Note that for normal k<strong>in</strong>ematics,<br />

where then | | | 0 | lead<strong>in</strong>g to a monotonic one-to-one mapp<strong>in</strong>g of the center-of-momentum<br />

angle and . However, for <strong>in</strong>verse k<strong>in</strong>ematics, where then | | | 0 | lead<strong>in</strong>g to two valued<br />

solutions at any fixed laboratory scatter<strong>in</strong>g angle .<br />

Billiard ball collisions are an especially simple example where the two masses³<br />

are identical ´ and the collision<br />

is essentially elastic. Then essentially =˜ =1, = <br />

2<br />

and = 1 2<br />

− , that is, the angle<br />

between the scattered billiard balls is 2 .<br />

Both normal and <strong>in</strong>verse k<strong>in</strong>ematics are illustrated <strong>in</strong> figure 1116 which shows the dependence of the<br />

projectile and target scatter<strong>in</strong>g angles <strong>in</strong> the laboratory frame as a function of center-of-momentum scatter<strong>in</strong>g<br />

angle for the Coulomb scatter<strong>in</strong>g of 104 Pd by 208 Pb, that is, for a mass ratio of 2:1. Both normal and<br />

<strong>in</strong>verse k<strong>in</strong>ematics are shown for the same bombard<strong>in</strong>g energy of 43 for elastic scatter<strong>in</strong>g and<br />

for <strong>in</strong>elastic scatter<strong>in</strong>g with a -value of −5.


278 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

Figure 11.17: Recoil energies, <strong>in</strong> , versus laboratory scatter<strong>in</strong>g angle, shown on the left for scatter<strong>in</strong>g<br />

of 447 104 Pd by 208 Pb with = −50, and shown on the right for scatter<strong>in</strong>g of 894 208 Pb<br />

on 104 Pd with = −50<br />

S<strong>in</strong>ce s<strong>in</strong>( − ) ≤ 1 then equation 11173 implies that ˜ s<strong>in</strong> ≤ 1 S<strong>in</strong>ce ˜ is always larger than<br />

or equal to unity there is a maximum scatter<strong>in</strong>g angle <strong>in</strong> the laboratory frame for the recoil<strong>in</strong>g target nucleus<br />

given by<br />

s<strong>in</strong> max = 1˜<br />

(11.176)<br />

For elastic scatter<strong>in</strong>g =s<strong>in</strong> −1 ( 1˜ )=90◦ s<strong>in</strong>ce ˜ =1for both 894 208 Pb bombard<strong>in</strong>g 104 Pd, and<br />

the <strong>in</strong>verse reaction us<strong>in</strong>g a 447 104 Pd beam scattered by a 208 Pb target. A -value of −5<br />

gives ˜ =1002808 which implies a maximum scatter<strong>in</strong>g angle of =8571 ◦ for both 894 208 Pb<br />

bombard<strong>in</strong>g 104 Pd, and the <strong>in</strong>verse reaction of a 447 104 Pd beam scattered by a 208 Pb target. As a<br />

consequence there are two solutions for for any allowed value of as illustrated <strong>in</strong> figure 1116.<br />

S<strong>in</strong>ce s<strong>in</strong>( − ) ≤ 1 then equation 11150 implies that s<strong>in</strong> ≤ 1 For a 447 104 Pd beam<br />

scattered by a 208 Pb target <br />

<br />

=050, thus =05 for elastic scatter<strong>in</strong>g which implies that there is no<br />

upper bound to . This leads to a one-to-one correspondence between and for normal k<strong>in</strong>ematics.<br />

In contrast, the projectile has a maximum scatter<strong>in</strong>g angle <strong>in</strong> the laboratory frame for <strong>in</strong>verse k<strong>in</strong>ematics<br />

s<strong>in</strong>ce <br />

<br />

=20 lead<strong>in</strong>g to an upper bound to given by<br />

s<strong>in</strong> max = 1 <br />

(11.177)<br />

For elastic scatter<strong>in</strong>g =2imply<strong>in</strong>g max =30 ◦ . In addition to hav<strong>in</strong>g a maximum value for , when<br />

1 also there are two solutions for for any allowed value of . For the example of 894 208 Pb<br />

bombard<strong>in</strong>g 178 Hf leads to a maximum projectile scatter<strong>in</strong>g angle of =300 ◦ for elastic scatter<strong>in</strong>g and<br />

=29907 ◦ for = −5<br />

K<strong>in</strong>etic energies<br />

The <strong>in</strong>itial total k<strong>in</strong>etic energy <strong>in</strong> the center-of-momentum frame is<br />

<br />

= (11.178)<br />

+ <br />

The f<strong>in</strong>al total k<strong>in</strong>etic energy <strong>in</strong> the center-of-momentum frame is<br />

<br />

<br />

= <br />

<br />

<br />

+ = ˜<br />

(11.179)<br />

+


11.13. TWO-BODY KINEMATICS 279<br />

In the laboratory frame the k<strong>in</strong>etic energies of the scattered projectile and recoil<strong>in</strong>g target nucleus are<br />

given by<br />

<br />

=<br />

<br />

=<br />

µ 2<br />

³ ´<br />

<br />

1+ 2 +2 cos ˜ (11.180)<br />

+ <br />

<br />

³<br />

<br />

( + ) 2 1+˜ 2 +2˜ cos ´ ˜ (11.181)<br />

where and are the center-of-mass scatter<strong>in</strong>g angles respectively for the scattered projectile and<br />

target nuclei.<br />

For the chosen <strong>in</strong>cident energies the normal and <strong>in</strong>verse reactions give the same center-of-momentum<br />

energy of 298 which is the energy available to the <strong>in</strong>teraction between the collid<strong>in</strong>g nuclei. However,<br />

the k<strong>in</strong>etic energy of the center-of-momentum is 447−298 = 149 for normal k<strong>in</strong>ematics and 894−298 =<br />

596 for <strong>in</strong>verse k<strong>in</strong>ematics. This trivial center-of-momentum k<strong>in</strong>etic energy does not contribute to the<br />

reaction. Note that <strong>in</strong>verse k<strong>in</strong>ematics focusses all the scattered nuclei <strong>in</strong>to the forward hemisphere which<br />

reduces the required solid angle for recoil-particle detection.<br />

Solid angles<br />

The laboratory-frame solid angles for the scattered projectile and target are taken to be and <br />

respectively, while the center-of-momentum solid angles are Ω and Ω respectively. The Jacobian relat<strong>in</strong>g<br />

the solid angles is<br />

Ã<br />

s<strong>in</strong> !2 <br />

<br />

=<br />

¯¯¯cos(<br />

Ω s<strong>in</strong> − ) ¯ (11.182)<br />

<br />

Ã<br />

s<strong>in</strong> !2 <br />

<br />

=<br />

¯¯¯cos(<br />

Ω s<strong>in</strong> − ) ¯ (11.183)<br />

<br />

These can be used to transform the calculated center-of-momentum differential cross sections to the<br />

laboratory frame for comparison with measured values. Note that relative to the center-of-momentum frame,<br />

the forward focuss<strong>in</strong>g <strong>in</strong>creases the observed differential cross sections <strong>in</strong> the forward laboratory frame and<br />

decreases them <strong>in</strong> the backward hemisphere.<br />

Exploitation of two-body k<strong>in</strong>ematics<br />

Comput<strong>in</strong>g the above non-trivial transform relations between the center-of-mass and laboratory coord<strong>in</strong>ate<br />

frames for two-body scatter<strong>in</strong>g is used extensively <strong>in</strong> many fields of physics. This discussion has assumed nonrelativistic<br />

two-body k<strong>in</strong>ematics. Relativistic two-body k<strong>in</strong>ematics encompasses non-relativistic k<strong>in</strong>ematics<br />

as discussed <strong>in</strong> chapter 174. Many computer codes are available that can be used for mak<strong>in</strong>g either nonrelativistic<br />

or relativistic transformations.<br />

It is stressed that the underly<strong>in</strong>g physics for two <strong>in</strong>teract<strong>in</strong>g bodies is identical irrespective of whether<br />

the reaction is observed <strong>in</strong> the center-of-mass or the laboratory coord<strong>in</strong>ate frames. That is, no new physics is<br />

<strong>in</strong>volved <strong>in</strong> the k<strong>in</strong>ematic transformation. However, the transformation between these frames can dramatically<br />

alter the angles and velocities of the observed scattered bodies which can be beneficial for experimental<br />

detection. For example, <strong>in</strong> heavy-ion nuclear physics the projectile and target nuclei can be <strong>in</strong>terchanged<br />

lead<strong>in</strong>g to very different velocities and scatter<strong>in</strong>g angles <strong>in</strong> the laboratory frame of reference. This can greatly<br />

facilitate identification and observation of the velocities vectors of the scattered nuclei. In high-energy physics<br />

it is advantageous to collide beams hav<strong>in</strong>g identical, but opposite, l<strong>in</strong>ear momentum vectors, s<strong>in</strong>ce then the<br />

laboratory frame is the center-of-mass frame, and the energy required to accelerate the collid<strong>in</strong>g bodies is<br />

m<strong>in</strong>imized.


280 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

11.14 Summary<br />

This chapter has focussed on the classical mechanics of bodies <strong>in</strong>teract<strong>in</strong>g via conservative, two-body, central<br />

<strong>in</strong>teractions. The follow<strong>in</strong>g are the ma<strong>in</strong> topics presented <strong>in</strong> this chapter.<br />

Equivalent one-body representation for two bodies <strong>in</strong>teract<strong>in</strong>g via a central <strong>in</strong>teraction The<br />

equivalent one-body representation of the motion of two bodies <strong>in</strong>teract<strong>in</strong>g via a two-body central <strong>in</strong>teraction<br />

greatly simplifies solution of the equations of motion. The position vectors r 1 and r 2 are expressed <strong>in</strong> terms<br />

of the center-of-mass vector R plus total mass = 1 + 2 while the position vector r plus associated<br />

reduced mass = 1 2<br />

1+ 2<br />

describe the relative motion of the two bodies <strong>in</strong> the center of mass. The total<br />

Lagrangian then separates <strong>in</strong>to two <strong>in</strong>dependent parts<br />

= 1 ¯¯¯Ṙ<br />

2 ¯<br />

¯2 + (1116)<br />

where the center-of-mass Lagrangian is<br />

= 1 2 |ṙ|2 − ()<br />

(1117)<br />

Equations 1110, and1111 can be used to derive the actual spatial trajectories of the two bodies expressed<br />

<strong>in</strong> terms of r 1 and r 2 from the relative equations of motion, written <strong>in</strong> terms of R and r for the equivalent<br />

one-body solution..<br />

Angular momentum Noether’s theorem shows that the angular momentum is conserved if only a sphericallysymmetric<br />

two-body central force acts between the <strong>in</strong>teract<strong>in</strong>g two bodies. The plane of motion is perpendicular<br />

to the angular momentum vector and thus the Lagrangian can be expressed <strong>in</strong> polar coord<strong>in</strong>ates<br />

as<br />

= 1 ³<br />

2 ˙ 2 + ˙2´ 2 − () (1122)<br />

Differential orbit equation of motion The B<strong>in</strong>et transformation = 1 <br />

allows the center-of-mass<br />

Lagrangian for a central force F =()ˆr to be used to express the differential orbit equation for the<br />

radial motion as<br />

2 <br />

2 + = − 1<br />

2 2 ( 1 )<br />

(1139)<br />

The Lagrangian, and the Hamiltonian all were used to derive the equations of motion for two bodies <strong>in</strong>teract<strong>in</strong>g<br />

via a two-body, conservative, central <strong>in</strong>teraction. The general features of the conservation of angular<br />

momentum and conservation of energy for a two-body, central potential were presented.<br />

Inverse-square, two-body, central force The <strong>in</strong>verse-square, two-body, central force is of pivotal importance<br />

<strong>in</strong> nature s<strong>in</strong>ce it is applies to both the gravitational force and the Coulomb force. The underly<strong>in</strong>g<br />

symmetries of the <strong>in</strong>verse-square, two-body, central <strong>in</strong>teraction, lead to conservation of angular momentum,<br />

conservation of energy, Gauss’s law, and that the two-body orbits follow closed, degenerate, orbits that are<br />

conic sections, for which the eccentricity vector is conserved. The radial dependence, relative to the force<br />

center ly<strong>in</strong>g at one focus of the conic section, is given by<br />

1<br />

= − 2 [1 + cos ( − 0)]<br />

(1158)<br />

where the orbit eccentricity equals<br />

s<br />

= 1+ 2 2<br />

2<br />

(1162)<br />

These lead to Kepler’s three laws of motion for two bodies <strong>in</strong> a bound orbit due to the attractive gravitational<br />

force for which = − 1 2 . The <strong>in</strong>verse-square law is special <strong>in</strong> that the eccentricity vector A is a third<br />

<strong>in</strong>variant of the motion, where<br />

A ≡ (p × L)+(ˆr)<br />

(1186)


11.14. SUMMARY 281<br />

The eccentricity vector unambiguously def<strong>in</strong>es the orientation and direction of the major axis of the elliptical<br />

orbit. The <strong>in</strong>variance of the eccentricity vector, and the existence of stable closed orbits, are manifestations<br />

of the dynamical 04 symmetry.<br />

Isotropic, harmonic, two-body, central force The isotropic, harmonic, two-body, central <strong>in</strong>teraction<br />

is of <strong>in</strong>terest s<strong>in</strong>ce, like the <strong>in</strong>verse-square law force, it leads to closed elliptical orbits described by<br />

⎛ Ã ! 1<br />

1<br />

2 = ⎝1+<br />

2 1+ 2 2<br />

<br />

<br />

2 cos 2( − <br />

<br />

0 )⎞<br />

⎠<br />

(11107)<br />

where the eccentricity is given by<br />

à ! 1<br />

1+ 2 2<br />

<br />

2 = 2<br />

2 − 2 (11108)<br />

The harmonic force orbits are dist<strong>in</strong>ctly different from those for the <strong>in</strong>verse-square law <strong>in</strong> that the force center<br />

is at the center of the ellipse, rather than at the focus for the <strong>in</strong>verse-square law force. This elliptical orbit<br />

is reflection symmetric for the harmonic force, but not for the <strong>in</strong>verse square force. The isotropic harmonic<br />

two-body force leads to <strong>in</strong>variance of the symmetry tensor, A 0 which is an <strong>in</strong>variant of the motion analogous<br />

to the eccentricity vector A. This leads to stable closed orbits, which are manifestations of the dynamical<br />

3 symmetry.<br />

Orbit stability Bertrand’s theorem states that only the <strong>in</strong>verse square law and the l<strong>in</strong>ear radial dependences<br />

of the central forces lead to stable closed bound orbits that do not precess. These are manifestation<br />

of the dynamical symmetries that occur for these two specific radial forms of two-body forces.<br />

The three-body problem The difficulties encountered <strong>in</strong> solv<strong>in</strong>g the equations of motion for three bodies,<br />

that are <strong>in</strong>teract<strong>in</strong>g via two-body central forces, was discussed. The three-body motion can <strong>in</strong>clude the<br />

existence of chaotic motion. It was shown that solution of the three-body problem is simplified if either the<br />

planar approximation, or the restricted three-body approximation, are applicable.<br />

Two-body scatter<strong>in</strong>g The total and differential two-body scatter<strong>in</strong>g cross sections were <strong>in</strong>troduced. It<br />

was shown that for the <strong>in</strong>verse-square law force there is a simple relation between the impact parameter <br />

and scatter<strong>in</strong>g angle given by<br />

=<br />

cot (11155)<br />

2 2<br />

This led to the solution for the differential scatter<strong>in</strong>g cross-section for Rutherford scatter<strong>in</strong>g due to the<br />

Coulomb <strong>in</strong>teraction.<br />

<br />

Ω = 1 µ 2 1<br />

4 2 s<strong>in</strong> 4 (11159)<br />

2<br />

This cross section assumes elastic scatter<strong>in</strong>g by a repulsive two-body <strong>in</strong>verse-square central force. For scatter<strong>in</strong>g<br />

of nuclei <strong>in</strong> the Coulomb potential the constant is given to be<br />

= 2<br />

4 <br />

(11160)<br />

Two-body k<strong>in</strong>ematics The transformation from the center-of-momentum frame to laboratory frames of<br />

reference was <strong>in</strong>troduced. Such transformations are used extensively <strong>in</strong> many fields of physics for theoretical<br />

modell<strong>in</strong>g of scatter<strong>in</strong>g, and for analysis of experiment data.


282 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

Problems<br />

1. Listed below are several statements concern<strong>in</strong>g central force motion. For each statement, give the reason for<br />

why the statement is true. If a statement is only true <strong>in</strong> certa<strong>in</strong> situations, then expla<strong>in</strong> when it holds and<br />

when it doesn’t. The system referred to below consists of mass 1 located at 1 and mass 2 located at 2 .<br />

(a) The potential energy of the system depends only on the difference 1 − 2 ,noton 1 and 2 separately.<br />

(b) The potential energy of the system depends only on the magnitude of 1 − 2 , not the direction.<br />

(c) It is possible to choose an <strong>in</strong>ertial reference frame <strong>in</strong> which the center of mass of the system is at rest.<br />

(d) The total energy of the system is conserved.<br />

(e) The total angular momentum of the system is conserved.<br />

2. A particle of mass moves <strong>in</strong> a potential () =− 0 −2 2 .<br />

(a) Given the constant , f<strong>in</strong>d an implicit equation for the radius of the circular orbit. A circular orbit at<br />

= is possible if<br />

µ <br />

=0<br />

<br />

¯¯¯¯=<br />

where is the effective potential.<br />

(b) What is the largest value of for which a circular orbit exists? What is the value of the effective potential<br />

at this critical orbit?<br />

3. A particle of mass is observed to move <strong>in</strong> a spiral orbit given by the equation = , where is a constant.<br />

Is it possible to have such an orbit <strong>in</strong> a central force field? If so, determ<strong>in</strong>e the form of the force function.<br />

4. The <strong>in</strong>teraction energy between two atoms of mass is given by the Lennard-Jones potential, () =<br />

£ ( 0 ) 12 − 2( 0 ) 6¤<br />

(a) Determ<strong>in</strong>e the Lagrangian of the system where 1 and 2 are the positions of the first and second mass,<br />

respectively.<br />

(b) Rewrite the Lagrangian as a one-body problem <strong>in</strong> which the center-of-mass is stationary.<br />

(c) Determ<strong>in</strong>e the equilibrium po<strong>in</strong>t and show that it is stable.<br />

(d) Determ<strong>in</strong>e the frequency of small oscillations about the stable po<strong>in</strong>t.<br />

5. Consider two bodies of mass <strong>in</strong> circular orbit of radius 0 2, attracted to each other by a force () ,where<br />

is the distance between the masses.<br />

(a) Determ<strong>in</strong>e the Lagrangian of the system <strong>in</strong> the center-of-mass frame (H<strong>in</strong>t: a one-body problem subject<br />

toacentralforce).<br />

(b) Determ<strong>in</strong>e the angular momentum. Is it conserved?<br />

(c) Determ<strong>in</strong>e the equation of motion <strong>in</strong> <strong>in</strong> terms of the angular momentum and |F()|.<br />

(d) Expand your result <strong>in</strong> (c) about an equilibrium radius 0 and show that the condition for stability<br />

is, 0 ( 0 )<br />

( 0) + 3 0<br />

0<br />

6. Consider two charges of equal magnitude connected by a spr<strong>in</strong>g of spr<strong>in</strong>g constant 0 <strong>in</strong> circular orbit. Can<br />

the charges oscillate about some equilibrium? If so, what condition must be satisfied?<br />

7. Consider a mass <strong>in</strong> orbit around a mass , which is subject to a force = − 2 ˆ ,where is the distance<br />

between the masses. Show that the eccentricity vector = × − ˆ is conserved.


11.14. SUMMARY 283<br />

8. Show that the areal velocity is constant for a particle mov<strong>in</strong>g under the <strong>in</strong>fluence of an attractive force given<br />

by () =−. Calculate the time averages of the k<strong>in</strong>etic and potential energies and compare with the the<br />

results of the virial theorem.<br />

9. Assume that the Earth’s orbit is circular and that the Sun’s mass suddenly decreases by a factor of two. (a)<br />

What orbit will the earth then have? (b) Will the Earth escape the solar system?<br />

10. Discuss the motion of a particle <strong>in</strong> a central <strong>in</strong>verse-square-law force field for a superimposed force whose<br />

magnitude is <strong>in</strong>versely proportional to the cube of the distance from the particle to force center; that is<br />

() =− 2 − 3<br />

(k, 0)<br />

Show that the motion is described by a precess<strong>in</strong>g ellipse. Consider the cases<br />

a) 2 , b) = 2 c) 2 <br />

where is the angular momentum and the reduced mass.<br />

11. A communications satellite is <strong>in</strong> a circular orbit around the earth at a radius and velocity . A rocket<br />

accidentally fires quite suddenly, giv<strong>in</strong>g the rocket an outward velocity <strong>in</strong> addition to its orig<strong>in</strong>al tangential<br />

velocity <br />

a) Calculate the ratio of the new energy and angular momentum to the old.<br />

b) Describe the subsequent motion of the satellite and plot () () the net effective potential, and ()<br />

after the rocket fires.<br />

12. Two identical po<strong>in</strong>t objects, each of mass are bound by a l<strong>in</strong>ear two-body force = − where is the<br />

vector distance between the two po<strong>in</strong>t objects. The two po<strong>in</strong>t objects each slide on a horizontal frictionless<br />

plane subject to a vertical gravitational field . The two-body system is free to translate, rotate and oscillate<br />

on the surface of the frictionless plane.<br />

(a) Derive the Lagrangian for the complete system <strong>in</strong>clud<strong>in</strong>g translation and relative motion.<br />

(b) Use Noether’s theorem to identify all constants of motion.<br />

(c) Use the Lagrangian to derive the equations of motion for the system.<br />

(d) Derive the generalized momenta and the correspond<strong>in</strong>g Hamiltonian.<br />

(e) Derive the period for small amplitude oscillations of the relative motion of the two masses.<br />

13. A bound b<strong>in</strong>ary star system comprises two spherical stars of mass 1 and 2 bound by their mutual gravitational<br />

attraction. Assume that the only force act<strong>in</strong>g on the stars is their mutual gravitation attraction and let<br />

be the <strong>in</strong>stantaneous separation distance between the centers of the two stars where is much larger than<br />

the sum of the radii of the stars.<br />

(a) Show that the two-body motion of the b<strong>in</strong>ary star system can be represented by an equivalent one-body<br />

system and derive the Lagrangian for this system.<br />

(b) Show that the motion for the equivalent one-body system <strong>in</strong> the center of mass frame lies entirely <strong>in</strong> a<br />

plane and derive the angle between the normal to the plane and the angular momentum vector.<br />

(c) Show whether is a constant of motion and whether it equals the total energy.<br />

(d) It is known that a solution to the equation of motion for the equivalent one-body orbit for this gravitational<br />

force has the form<br />

1<br />

= − [1 + cos ]<br />

2 and that the angular momentum is a constant of motion = .<br />

force lead<strong>in</strong>g to this bound orbit is<br />

F = 2ˆr<br />

where must be negative.<br />

Use these to prove that the attractive


284 CHAPTER 11. CONSERVATIVE TWO-BODY CENTRAL FORCES<br />

14. When perform<strong>in</strong>g the Rutherford experiment, Gieger and Marsden scattered 77 4 He particles (alpha<br />

particles) from 238 U at a scatter<strong>in</strong>g angle <strong>in</strong> the laboratory frame of =90 0 . Derive the follow<strong>in</strong>g observables<br />

as measured <strong>in</strong> the laboratory frame.<br />

(a) The recoil scatter<strong>in</strong>g angle of the 238 U <strong>in</strong> the laboratory frame.<br />

(b) The scatter<strong>in</strong>g angles of the 4 He and 238 U <strong>in</strong> the center-of-mass frame<br />

(c) The k<strong>in</strong>etic energies of the 4 He and 238 U <strong>in</strong> the laboratory frame<br />

(d) The impact parameter<br />

(e) The distance of closest approach m<strong>in</strong>


Chapter 12<br />

Non-<strong>in</strong>ertial reference frames<br />

12.1 Introduction<br />

Newton’s Laws of motion apply only to <strong>in</strong>ertial frames of reference. Inertial frames of reference make it<br />

possible to use either Newton’s laws of motion, or Lagrangian, or Hamiltonian mechanics, to develop the<br />

necessary equations of motion. There are certa<strong>in</strong> situations where it is much more convenient to treat the<br />

motion <strong>in</strong> a non-<strong>in</strong>ertial frame of reference. Examples are motion <strong>in</strong> frames of reference undergo<strong>in</strong>g translational<br />

acceleration, rotat<strong>in</strong>g frames of reference, or frames undergo<strong>in</strong>g both translational and rotational<br />

motion. This chapter will analyze the behavior of dynamical systems <strong>in</strong> accelerated frames of reference,<br />

especially rotat<strong>in</strong>g frames such as on the surface of the Earth. Newtonian mechanics, as well as the Lagrangian<br />

and Hamiltonian approaches, will be used to handle motion <strong>in</strong> non-<strong>in</strong>ertial reference frames by<br />

<strong>in</strong>troduc<strong>in</strong>g extra <strong>in</strong>ertial forces that correct for the fact that the motion is be<strong>in</strong>g treated with respect to a<br />

non-<strong>in</strong>ertial reference frame. These <strong>in</strong>ertial forces are often called fictitious even though they appear real <strong>in</strong><br />

the non-<strong>in</strong>ertial frame. The underly<strong>in</strong>g reasons for each of the <strong>in</strong>ertial forces will be discussed followed by a<br />

presentation of important applications.<br />

12.2 Translational acceleration of a reference frame<br />

Consider an <strong>in</strong>ertial system ( ) which is fixed<br />

<strong>in</strong> space, and a non-<strong>in</strong>ertial system ( 0 0 0 ) that<br />

is mov<strong>in</strong>g <strong>in</strong> a direction relative to the fixed frame such as<br />

to ma<strong>in</strong>ta<strong>in</strong> constant orientations of the axes relative to the<br />

fixed frame, as illustrated <strong>in</strong> figure 121. The fixed frame is<br />

designated to be the unprimed frame and, to avoid confusion<br />

the subscript is attached to the fixed coord<strong>in</strong>ates<br />

taken with respect to the fixed coord<strong>in</strong>ate frame. Similarly,<br />

the translat<strong>in</strong>g reference frame, which is undergo<strong>in</strong>g translational<br />

acceleration, has the subscript attached to the<br />

coord<strong>in</strong>ates taken with respect to the translat<strong>in</strong>g frame of<br />

reference. Newton’s Laws of motion are obeyed only <strong>in</strong> the<br />

<strong>in</strong>ertial (unprimed) reference frame. The respective position<br />

vectors are related by<br />

r = R +r 0 (12.1)<br />

where r is the vector relative to the fixed frame, r 0 is<br />

the vector relative to the translationally accelerat<strong>in</strong>g frame<br />

and R is the vector from the orig<strong>in</strong> of the fixed frame to<br />

the orig<strong>in</strong> of the accelerat<strong>in</strong>g frame. Differentiat<strong>in</strong>g equation<br />

121 gives the velocity vector relation<br />

v = V +v 0 (12.2)<br />

285<br />

Figure 12.1: Inertial reference frame (unprimed),<br />

and translational accelerat<strong>in</strong>g frame<br />

(primed).


286 CHAPTER 12. NON-INERTIAL REFERENCE FRAMES<br />

where v = r <br />

<br />

v 0 = r0 <br />

<br />

where a = 2 r <br />

<br />

a 0 2 = 2 r 0 <br />

<br />

and A 2<br />

= 2 R <br />

<br />

2<br />

In the fixed frame, Newton’s laws give that<br />

and V = R <br />

<br />

. Similarly the acceleration vector relation is<br />

a = A +a 0 (12.3)<br />

F = a (12.4)<br />

The force <strong>in</strong> the fixed frame can be separated <strong>in</strong>to two terms, the acceleration of the accelerat<strong>in</strong>g frame of<br />

reference A plus the acceleration with respect to the accelerat<strong>in</strong>g frame a 0 .<br />

Relative to the accelerat<strong>in</strong>g reference frame the acceleration is given by<br />

F = A +a 0 (12.5)<br />

a 0 = F − A (12.6)<br />

The accelerat<strong>in</strong>g frame of reference can exploit Newton’s Laws of motion us<strong>in</strong>g an effective translational<br />

force F 0 ≡ F − A The additional −A term is called an <strong>in</strong>ertial force; it can be altered by<br />

choos<strong>in</strong>g a different non-<strong>in</strong>ertial frame of reference, that is, it is dependent on the frame of reference <strong>in</strong> which<br />

the observer is situated.<br />

12.3 Rotat<strong>in</strong>g reference frame<br />

Consider a rotat<strong>in</strong>g frame of reference which will be designated as the double-primed (rotat<strong>in</strong>g) frame<br />

to differentiate it from the non-rotat<strong>in</strong>g primed (mov<strong>in</strong>g) frame, s<strong>in</strong>ce both of which may be undergo<strong>in</strong>g<br />

translational acceleration relative to the <strong>in</strong>ertial fixed unprimed frame as described above.<br />

12.3.1 Spatial time derivatives <strong>in</strong> a rotat<strong>in</strong>g, non-translat<strong>in</strong>g, reference frame<br />

For simplicity assume that R = V =0 that is, the<br />

primed reference frame is stationary and identical to the fixed<br />

stationary unprimed frame. The double-primed (rotat<strong>in</strong>g)<br />

frame is a non-<strong>in</strong>ertial frame rotat<strong>in</strong>g with respect to the<br />

orig<strong>in</strong> of the fixed primed frame. Appendix 23 shows that<br />

an <strong>in</strong>f<strong>in</strong>itessimal rotation about an <strong>in</strong>stantaneous axis of<br />

rotation leads to an <strong>in</strong>f<strong>in</strong>itessimal displacement r where<br />

r = θ × r 0 (12.7)<br />

Consider that dur<strong>in</strong>g a time the position vector <strong>in</strong> the fixed<br />

primed reference frame moves by an arbitrary <strong>in</strong>f<strong>in</strong>itessimal<br />

distance r 0 As illustrated <strong>in</strong> figure 122, this<strong>in</strong>f<strong>in</strong>itessimal<br />

distance <strong>in</strong> the primed non-rotat<strong>in</strong>g frame can be split<br />

<strong>in</strong>to two parts:<br />

a) r = θ×r 0 which is due to rotation of the rotat<strong>in</strong>g<br />

frame with respect to the translat<strong>in</strong>g primed frame.<br />

b) (r 00<br />

) which is the motion with respect to the rotat<strong>in</strong>g<br />

(double-primed) frame.<br />

That is, the motion has been arbitrarily divided <strong>in</strong>to<br />

a part that is due to the rotation of the double-primed<br />

frame, plus the vector displacement measured <strong>in</strong> this rotat<strong>in</strong>g<br />

(double-primed) frame. It is always possible to make such a<br />

decomposition of the displacement as long as the vector sum<br />

can be written as<br />

r 0 = r 00<br />

+ θ × r 0 (12.8)<br />

Figure 12.2: Inf<strong>in</strong>itessimal displacement <strong>in</strong><br />

thenonrotat<strong>in</strong>gprimedframeand<strong>in</strong>therotat<strong>in</strong>g<br />

double-primed reference frame frame.


12.3. ROTATING REFERENCE FRAME 287<br />

S<strong>in</strong>ce θ = ω thenthetimedifferential of the displacement, equation 128, canbewrittenas<br />

µ µ <br />

r<br />

0 r<br />

00<br />

= + ω × r 0 (12.9)<br />

<br />

<br />

<br />

<br />

³ ´<br />

r<br />

The important conclusion is that a velocity measured <strong>in</strong> a non-rotat<strong>in</strong>g reference frame<br />

0<br />

<br />

can be<br />

³ ´<br />

<br />

r<br />

expressed as the sum of the velocity<br />

00<br />

<br />

measured relative to a rotat<strong>in</strong>g frame, plus the term ω ×r0 <br />

which accounts for the rotation of the frame. The division of the r 0 vector <strong>in</strong>to two parts, a part due to<br />

rotation of the frame plus a part with respect to the rotat<strong>in</strong>g frame, is valid for any vector as shown below.<br />

12.3.2 General vector <strong>in</strong> a rotat<strong>in</strong>g, non-translat<strong>in</strong>g, reference frame<br />

Consider an arbitrary vector G which can be expressed <strong>in</strong> terms of components along the three unit vector<br />

basis ê <br />

<strong>in</strong> the fixed <strong>in</strong>ertial frame as<br />

3X<br />

G = <br />

ê <br />

(12.10)<br />

=1<br />

Neglect<strong>in</strong>g translational motion, then it can be expressed <strong>in</strong> terms of the three unit vectors <strong>in</strong> the non-<strong>in</strong>ertial<br />

rotat<strong>in</strong>g frame unit vector basis ê <br />

as<br />

3X<br />

G = ( ) ê <br />

(12.11)<br />

S<strong>in</strong>ce the unit basis vectors ê <br />

<br />

=1<br />

are constant <strong>in</strong> the rotat<strong>in</strong>g frame, that is,<br />

µ ê<br />

<br />

<br />

=0 (12.12)<br />

<br />

then the time derivatives of G <strong>in</strong> the rotat<strong>in</strong>g coord<strong>in</strong>ate system ê <br />

can be written as<br />

µ G<br />

3X<br />

µ <br />

=<br />

ê <br />

(12.13)<br />

<br />

<br />

<br />

=1<br />

<br />

The <strong>in</strong>ertial-frame time derivative taken with components along the rotat<strong>in</strong>g coord<strong>in</strong>ate basis ê <br />

,equation<br />

1211, is<br />

µ G<br />

3X<br />

µ <br />

3X<br />

=<br />

ê <br />

ê <br />

<br />

+ ( )<br />

<br />

<br />

<br />

<br />

(12.14)<br />

=1<br />

<br />

<br />

=1<br />

Substitute the unit vector ê for r 0 <strong>in</strong> equation 129 plus us<strong>in</strong>g equation 1212 gives that<br />

µ ê<br />

<br />

= ω × ê (12.15)<br />

<br />

<br />

Substitute this <strong>in</strong>to the second term of equation 1214 gives<br />

µ µ <br />

G G<br />

= + ω × G (12.16)<br />

<br />

<br />

<br />

<br />

This important identity relates the time derivatives of any vector expressed <strong>in</strong> both the <strong>in</strong>ertial frame and<br />

the rotat<strong>in</strong>g non-<strong>in</strong>ertial frame bases. Note that the ω × G term orig<strong>in</strong>ates from the fact that the unit<br />

basis vectors of the rotat<strong>in</strong>g reference frame are time dependent with respect to the non-rotat<strong>in</strong>g frame basis<br />

vectors as given by equation (1215). Equation (1216) is used extensively for problems <strong>in</strong>volv<strong>in</strong>g rotat<strong>in</strong>g<br />

frames. For example, for the special case where G = r 0 ,thenequation(1216) relates the velocity vectors <strong>in</strong><br />

the fixed and rotat<strong>in</strong>g frames as given <strong>in</strong> equation (129).<br />

Another example is the vector ˙ω<br />

µ µ <br />

µ <br />

ω ω<br />

ω<br />

˙ω = = + ω × ω = = ˙ω (12.17)<br />

<br />

<br />

<br />

<br />

<br />

<br />

That is, the angular acceleration ˙ has the same value <strong>in</strong> both the fixed and rotat<strong>in</strong>g frames of reference.


288 CHAPTER 12. NON-INERTIAL REFERENCE FRAMES<br />

12.4 Reference frame undergo<strong>in</strong>g rotation plus translation<br />

Consider the case where the system is accelerat<strong>in</strong>g <strong>in</strong> translation as well as rotat<strong>in</strong>g, that is, the primed<br />

frame is the non-rotat<strong>in</strong>g translat<strong>in</strong>g frame. The position vector r is taken with respect to the <strong>in</strong>ertial<br />

fixedunprimedframewhichcanbewritten<strong>in</strong>termsofthefixed unit basis vectors ( b i b j b k ) This r <br />

vector can be written as the vector sum of the translational motion R of the orig<strong>in</strong> of the rotat<strong>in</strong>g system<br />

with respect to the fixed frame, plus the position r 0 with respect to this translat<strong>in</strong>g primed frame basis<br />

Thetimedifferential is<br />

µ r<br />

=<br />

<br />

<br />

r = R + r 0 (12.18)<br />

µ R<br />

+<br />

<br />

<br />

µ r<br />

0<br />

<br />

<br />

(12.19)<br />

The vector r 0 is the position ³ with respect to the translat<strong>in</strong>g frame of reference which can be expressed <strong>in</strong><br />

terms of the unit vectors bi 0 j b0 k b ´<br />

0 .<br />

Equation 1219 takes <strong>in</strong>to account the translational motion of the mov<strong>in</strong>g primed frame basis. Now,<br />

assum<strong>in</strong>g that the double primed frame rotates about the orig<strong>in</strong> of the mov<strong>in</strong>g primed frame, then the net<br />

displacement with respect to the orig<strong>in</strong>al <strong>in</strong>ertial frame basis can be comb<strong>in</strong>ed with equation 129 lead<strong>in</strong>g to<br />

the relation<br />

µ µ µ <br />

r R r<br />

00<br />

= + + ω × r 0 (12.20)<br />

<br />

<br />

<br />

<br />

<br />

<br />

Here the double-primed frame ³ is both rotat<strong>in</strong>g and translat<strong>in</strong>g. Vectors <strong>in</strong> this frame are expressed <strong>in</strong> terms<br />

of the unit basis vectors bi 00 j b00 k c ´<br />

00 <br />

Expressed as velocities, equation 1220 can be written as<br />

v = V + v 00 + ω × r 0 (12.21)<br />

where:<br />

v is the velocity measured with respect to the <strong>in</strong>ertial (unprimed) frame basis.<br />

V is the velocity of the orig<strong>in</strong> of the non-<strong>in</strong>ertial translat<strong>in</strong>g (primed) frame basis with respect to the<br />

orig<strong>in</strong> of the <strong>in</strong>ertial (unprimed) frame basis.<br />

v 00 is the velocity of the particle with respect to the non-<strong>in</strong>ertial rotat<strong>in</strong>g (double-primed) frame basis<br />

the orig<strong>in</strong> of which is both translat<strong>in</strong>g and rotat<strong>in</strong>g.<br />

ω × r 0 is the motion of the rotat<strong>in</strong>g (double-primed) frame with respect to the l<strong>in</strong>early-translat<strong>in</strong>g<br />

(primed) frame basis.<br />

Thus this relation takes <strong>in</strong>to account both the translational velocity plus rotation of the reference coord<strong>in</strong>ate<br />

frame basis vectors.<br />

12.5 Newton’s law of motion <strong>in</strong> a non-<strong>in</strong>ertial frame<br />

The acceleration of the system <strong>in</strong> the rotat<strong>in</strong>g <strong>in</strong>ertial frame can be derived by differentiat<strong>in</strong>g the general<br />

velocity relation for v equation 1221 <strong>in</strong> the fixed frame basis which gives<br />

µ µ µ µ <br />

µ <br />

v<br />

V v<br />

00<br />

ω<br />

r<br />

a =<br />

=<br />

+<br />

+<br />

× r 0 0<br />

+ ω × <br />

(12.22)<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Now we wish to use the general transformation to a rotat<strong>in</strong>g frame basis which requires <strong>in</strong>clusion of the time<br />

dependence of the unit vectors <strong>in</strong> the rotat<strong>in</strong>g frame, that is,<br />

µ µ <br />

v<br />

00<br />

<br />

v<br />

00<br />

<br />

=<br />

+ ω × v 00<br />

(12.23)<br />

<br />

<br />

<br />

<br />

µ <br />

µ <br />

ω<br />

ω<br />

× r 0 =<br />

× r 0 (12.24)<br />

<br />

<br />

<br />

<br />

µ r<br />

0<br />

ω × <br />

= ω × v 00 + ω × (ω × r 0<br />

<br />

) (12.25)


12.6. LAGRANGIAN MECHANICS IN A NON-INERTIAL FRAME 289<br />

Us<strong>in</strong>g equations 1223 1224 1225 gives<br />

a = A + a 00<br />

+2ω × v 00 + ω × (ω × r 0 )+ ˙ω × r 0 (12.26)<br />

µ<br />

where the acceleration <strong>in</strong> the rotat<strong>in</strong>g frame is a 00 v 00<br />

<br />

=<br />

<br />

A is with respect to the fixed frame.<br />

Newton’s laws of motion are obeyed <strong>in</strong> the <strong>in</strong>ertial frame, that is<br />

<br />

<br />

while the velocity is v 00<br />

=<br />

µ<br />

r 00<br />

<br />

<br />

F = a = (A + a 00<br />

+2ω × v 00<br />

+ ω × (ω × r 0 )+ ˙ω × r 0 ) (12.27)<br />

In the double-primed frame, which may be both rotat<strong>in</strong>g and accelerat<strong>in</strong>g <strong>in</strong> translation, one can ascribe an<br />

effective force F <br />

that obeys an effective Newton’s law for the acceleration a 00<br />

<strong>in</strong> the rotat<strong>in</strong>g frame<br />

F <br />

= a 00<br />

= F − (A +2ω × v 00<br />

+ ω × (ω × r 0 )+ ˙ω × r 0 ) (12.28)<br />

Note that the effective force F <br />

comprises the physical force F m<strong>in</strong>us four non-<strong>in</strong>ertial forces that are<br />

<strong>in</strong>troduced to correct for the fact that the rotat<strong>in</strong>g reference frame is a non-<strong>in</strong>ertial frame.<br />

12.6 Lagrangian mechanics <strong>in</strong> a non-<strong>in</strong>ertial frame<br />

The above derivation of the equations of motion <strong>in</strong> the rotat<strong>in</strong>g frame is based on Newtonian mechanics.<br />

Lagrangian mechanics provides another derivation of these equations of motion for a rotat<strong>in</strong>g frame of<br />

reference by exploit<strong>in</strong>g the fact that the Lagrangian is a scalar which is frame <strong>in</strong>dependent, that is, it is<br />

<strong>in</strong>variant to rotation of the frame of reference.<br />

The Lagrangian <strong>in</strong> any frame is given by<br />

= 1 v · v − () (12.29)<br />

2<br />

The scalar product v · v is the same <strong>in</strong> any rotated frame and can be evaluated <strong>in</strong> terms of the rotat<strong>in</strong>g<br />

frame variables us<strong>in</strong>g the same decomposition of the translational plus rotational motion as used previously<br />

and given <strong>in</strong> equation 1221<br />

Equation (1221) decomposes the velocity <strong>in</strong> the fixed <strong>in</strong>ertial frame v <strong>in</strong>to four vector terms, the<br />

translational velocity V of the translat<strong>in</strong>g frame, the velocity <strong>in</strong> the rotat<strong>in</strong>g-translat<strong>in</strong>g frame v 00 and<br />

rotational velocity (ω × r 0 ). Us<strong>in</strong>g equations 1229 and 1221 plus appendix equation 21 for the triple<br />

products, gives that the Lagrangian evaluated us<strong>in</strong>g v ·v equals<br />

= 1 h<br />

2 V ·V +v 00<br />

·v 00<br />

+2V ·v 00<br />

+2V · (ω × r 0 )+2v 00 · (ω × r 0 )+(ω × r 0 ) 2i −()<br />

(12.30)<br />

This can be used to derive the canonical momentum <strong>in</strong> the rotat<strong>in</strong>g frame<br />

p 00<br />

=<br />

<br />

v<br />

00 = [V +v 00<br />

+ ω × r 0 ] (12.31)<br />

The Lagrange equations can be used to derive the equations of motion <strong>in</strong> terms of the variables evaluated<br />

<strong>in</strong> the rotat<strong>in</strong>g reference frame. The required Lagrange derivatives are<br />

<br />

<br />

<br />

v 00<br />

<br />

<br />

<br />

and<br />

= [A +a 00<br />

+(ω × v 00<br />

)+(˙ω × r 0 )] <br />

(12.32)<br />

and<br />

<br />

r 0 = − [(ω × V ) − (ω × v) 00 − ω × (ω × r 0 )] <br />

− ∇ (12.33)<br />

where the scalar triple product, equation 21 has been used. Thus the Lagrange equations give for the<br />

rotat<strong>in</strong>g frame basis that<br />

a 00<br />

= −∇ − [A +(ω × V )+2(ω × v 00<br />

)+ω × (ω × r 0 )+(˙ω × r 0 )] (12.34)


290 CHAPTER 12. NON-INERTIAL REFERENCE FRAMES<br />

The external force is identified as F = −∇. Equation 1216 can be used to transform between the<br />

fixed and the rotat<strong>in</strong>g bases.<br />

i<br />

A =<br />

hA +(ω × V) <br />

(12.35)<br />

This leads to an effective force <strong>in</strong> the non-<strong>in</strong>ertial translat<strong>in</strong>g plus rotat<strong>in</strong>g frame that corresponds to an<br />

effective Newtonian force of<br />

F <br />

= a 00<br />

= F − [A +2ω × v 00 + ω × (ω × r 0 )+(˙ω × r 0 )] (12.36)<br />

where A is expressed <strong>in</strong> the fixed frame. The derivation of equation 1236 us<strong>in</strong>g Lagrangian mechanics,<br />

confirms the identical formula 1229 derived us<strong>in</strong>g Newtonian mechanics.<br />

The four correction terms for the non-<strong>in</strong>ertial frame basis correspond to the follow<strong>in</strong>g effective forces.<br />

Translational acceleration: F <br />

= −A is the usual <strong>in</strong>ertial force experienced <strong>in</strong> a l<strong>in</strong>early accelerat<strong>in</strong>g<br />

frame of reference, and where A is with respect to the fixed frame .<br />

Coriolis force; F <br />

= −2ω × v 00 This is a new type of <strong>in</strong>ertial force that is present only when a<br />

particle is mov<strong>in</strong>g <strong>in</strong> the rotat<strong>in</strong>g frame. This force is proportional to the velocity <strong>in</strong> the rotat<strong>in</strong>g frame and<br />

is <strong>in</strong>dependent of the position <strong>in</strong> the rotat<strong>in</strong>g frame<br />

Centrifugal force: F <br />

<br />

= −ω × (ω × r 0 ) This is due to the centripetal acceleration of the particle<br />

ow<strong>in</strong>g to the rotation of the mov<strong>in</strong>g axis about the axis of rotation.<br />

Transverse (azimuthal) force: F <br />

= − ˙ω × r 0 This is a straightforward term due to acceleration of<br />

the particle due to the angular acceleration of the rotat<strong>in</strong>g axes.<br />

The above <strong>in</strong>ertial forces are correction terms aris<strong>in</strong>g from try<strong>in</strong>g to extend Newton’s laws of motion to<br />

a non-<strong>in</strong>ertial frame <strong>in</strong>volv<strong>in</strong>g both translation and rotation. These correction forces are often referred to as<br />

“fictitious” forces. However, these non-<strong>in</strong>ertial forces are very real when located <strong>in</strong> the non-<strong>in</strong>ertial frame.<br />

S<strong>in</strong>ce the centrifugal and Coriolis terms are unusual they are discussed below.<br />

12.7 Centrifugal force<br />

The centrifugal force was def<strong>in</strong>ed as<br />

F = −ω × (ω × r 0 ) (12.37)<br />

<br />

Note that<br />

ω · F =0 (12.38)<br />

therefore the centrifugal force is perpendicular to the axis of<br />

rotation.<br />

Us<strong>in</strong>g the vector identity, equation 24 allows the centrifugal<br />

force to be written as<br />

F = − £ (ω · r 0 ) ω − 2 r 0 ¤<br />

(12.39)<br />

For the case where the radius r 0 is perpendicular to ω then ω·r 0 =<br />

0 and thus for this special case<br />

ṙ<br />

F = 2 r 0 (12.40)<br />

The centrifugal force is experienced when rid<strong>in</strong>g <strong>in</strong> a car<br />

driven rapidly around a bend. The passenger experiences an apparent<br />

centrifugal (center flee<strong>in</strong>g) force that thrusts them to the<br />

outside of the bend relative to the <strong>in</strong>side of the turn<strong>in</strong>g car. In<br />

reality, relative to the fixed <strong>in</strong>ertial frame, i.e. the road, the friction<br />

between the car tires and the road is chang<strong>in</strong>g the direction<br />

of the car towards the <strong>in</strong>side of the bend and the car seat is caus<strong>in</strong>g<br />

the centripetal (center seek<strong>in</strong>g) acceleration of the passenger.<br />

A bucket of water attached to a rope can be swung around <strong>in</strong> a<br />

vertical plane without spill<strong>in</strong>g any water if the centrifugal force<br />

exceeds the gravitation force at the top of the trajectory.<br />

O<br />

Figure 12.3: Centrifugal force.


12.8. CORIOLIS FORCE 291<br />

12.8 Coriolis force<br />

The Coriolis force was def<strong>in</strong>ed to be<br />

F = −2ω × v 00 (12.41)<br />

where v 00 is the velocity measured <strong>in</strong> the rotat<strong>in</strong>g<br />

(double-primed) frame. The Coriolis<br />

force is an <strong>in</strong>terest<strong>in</strong>g force; it is perpendicular<br />

to both the axis of rotation and the velocity<br />

vector <strong>in</strong> the rotat<strong>in</strong>g frame, that is, it<br />

is analogous to the v × B Lorentz magnetic<br />

force .<br />

The understand<strong>in</strong>g of the Coriolis effect<br />

is facilitated by consider<strong>in</strong>g the physics of a<br />

hockey puck slid<strong>in</strong>g on a rotat<strong>in</strong>g frictionless<br />

Figure 12.4: Free-force motion of a hockey puck slid<strong>in</strong>g on<br />

table. Assume that the table rotates with<br />

a rotat<strong>in</strong>g frictionless table of radius that is rotat<strong>in</strong>g with<br />

constant angular frequency ω = k b about<br />

constant angular frequency out of the page.<br />

the axis. For this system the orig<strong>in</strong> of the<br />

rotat<strong>in</strong>g system is fixed, and the angular frequency is constant, thus A and ˙ ×r 0 are zero. Also it is assumed<br />

that there are no external forces act<strong>in</strong>g on the hockey puck, thus the net acceleration of the puck slid<strong>in</strong>g on<br />

the table, as seen <strong>in</strong> the rotat<strong>in</strong>g frame, simplifies to<br />

a 00<br />

= −2ω × v 00 − ω × (ω × r 0 ) =−2ˆk × v 00 + 2 r 0 (12.42)<br />

The centrifugal acceleration + 2 r 0 is radially outwards while the Coriolis acceleration −2k b × v 00 is to<br />

the right. Integration of the equations of motion can be used to calculate the trajectories <strong>in</strong> the rotat<strong>in</strong>g<br />

frame of reference.<br />

Figure 124 illustrates trajectories of the hockey puck <strong>in</strong> the rotat<strong>in</strong>g reference frame when no external<br />

forces are act<strong>in</strong>g, that is, <strong>in</strong> the <strong>in</strong>ertial frame the puck moves <strong>in</strong> a straight l<strong>in</strong>e with constant velocity v 0 .<br />

In the rotat<strong>in</strong>g reference frame the Coriolis force accelerates the puck to the right lead<strong>in</strong>g to trajectories<br />

that exhibit spiral motion. The apparent complicated trajectories are a result of the observer be<strong>in</strong>g <strong>in</strong> the<br />

rotat<strong>in</strong>g frame for which that the straight <strong>in</strong>ertial-frame trajectories of the mov<strong>in</strong>g puck exhibit a spirall<strong>in</strong>g<br />

trajectory <strong>in</strong> the rotat<strong>in</strong>g-frame.<br />

The Coriolis force is the reason that w<strong>in</strong>ds circulate <strong>in</strong> an anticlockwise direction about low-pressure<br />

regions <strong>in</strong> the Earth’s northern hemisphere. It also has important consequences <strong>in</strong> many activities on earth<br />

such as ballet danc<strong>in</strong>g, ice skat<strong>in</strong>g, acrobatics, nuclear and molecular rotation, and the motion of missiles.<br />

12.1 Example: Accelerat<strong>in</strong>g spr<strong>in</strong>g plane pendulum<br />

Comparison of the relative merits of us<strong>in</strong>g a non-<strong>in</strong>ertial frame versus an <strong>in</strong>ertial frame is given by a<br />

spr<strong>in</strong>g pendulum attached to an accelerat<strong>in</strong>g fulcrum. As shown <strong>in</strong> the figure, the spr<strong>in</strong>g pendulum comprises<br />

amass attached to a massless spr<strong>in</strong>g that has a rest length 0 and spr<strong>in</strong>g constant . The system is<br />

<strong>in</strong> a vertical gravitational field and the fulcrum of the pendulum is accelerat<strong>in</strong>g vertically upwards with a<br />

constant acceleration . Assume that the spr<strong>in</strong>g pendulum oscillates only <strong>in</strong> the vertical plane.<br />

Inertial frame:<br />

This problem can be solved <strong>in</strong> the fixed <strong>in</strong>ertial coord<strong>in</strong>ate system with coord<strong>in</strong>ates ( ). These coord<strong>in</strong>ates,<br />

and their time derivatives, are given <strong>in</strong> terms of and by<br />

Thus<br />

= s<strong>in</strong> ˙ = ˙ s<strong>in</strong> + ˙ cos <br />

= − cos + 1 2 2 ˙ = ˙ s<strong>in</strong> − ˙ cos + <br />

= 1 2 ¡ ˙ 2 + ˙ 2¢ − − 1 2 ( − 0) 2<br />

= 1 h<br />

³<br />

2 ˙ 2 + 2 ˙2 + 2 2 +2 ˙<br />

´i<br />

s<strong>in</strong> − ˙ cos + <br />

µ cos − 1 <br />

2 2 − 1 2 ( − 0) 2


292 CHAPTER 12. NON-INERTIAL REFERENCE FRAMES<br />

The Lagrange equations of motion are given by<br />

Λ =0<br />

¨ − ˙ 2 − ( + )cos + ( − 0)=0<br />

Λ =0<br />

The generalized momenta are<br />

¨ + 2 ˙ ˙ +<br />

( + )<br />

<br />

s<strong>in</strong> =0<br />

= = ˙ − cos <br />

˙<br />

= <br />

˙ = 2 ˙ + s<strong>in</strong> <br />

These lead to the correspond<strong>in</strong>g velocities of<br />

and thus the Hamiltonian is given by<br />

˙ = <br />

+ cos <br />

<br />

˙ = s<strong>in</strong> <br />

−<br />

2 <br />

= ˙ + ˙ − <br />

= 2 <br />

2 + <br />

2 2 − <br />

s<strong>in</strong> + cos + 1 2 ( − 0) 2 + 1 2 2 − cos <br />

The Hamilton equations of motion give that<br />

˙ = = <br />

+ cos <br />

<br />

˙ = =<br />

s<strong>in</strong> <br />

−<br />

2 <br />

These radial and angular velocities are the same as obta<strong>in</strong>ed us<strong>in</strong>g Lagrangian mechanics.<br />

The Hamilton equations for ˙ and ˙ are given by<br />

˙ = − <br />

= − <br />

2 s<strong>in</strong> − ( − 0 )+ cos +<br />

Similarly<br />

˙ = − <br />

= <br />

cos + s<strong>in</strong> − s<strong>in</strong> <br />

The transformation equations relat<strong>in</strong>g the generalized coord<strong>in</strong>ates aretimedependentsotheHamiltonian<br />

does not equal the total energy . In addition neither the Lagrangian nor the Hamiltonian are<br />

conserved s<strong>in</strong>ce they both are time dependent. The fact that the Hamiltonian is not conserved is obvious s<strong>in</strong>ce<br />

the whole system is accelerat<strong>in</strong>g upwards lead<strong>in</strong>g to <strong>in</strong>creas<strong>in</strong>g k<strong>in</strong>etic and potential energies. Moreover, the<br />

time derivative of the angular momentum ˙ is non-zero so the angular momentum is not conserved.<br />

Non-<strong>in</strong>ertial fulcrum frame:<br />

This system also can be addressed <strong>in</strong> the accelerat<strong>in</strong>g non-<strong>in</strong>ertial fulcrum frame of reference which is<br />

fixed to the fulcrum of the spr<strong>in</strong>g of the pendulum. In this non-<strong>in</strong>ertial frame of reference, the acceleration<br />

of the frame can be taken <strong>in</strong>to account us<strong>in</strong>g an effective acceleration which is added to the gravitational<br />

force; that is, is replaced by an effective gravitational force ( + ). Then the Lagrangian <strong>in</strong> the fulcrum<br />

frame simplifies to<br />

= 1 2 ˙2 + 2 1<br />

˙2 + ( + )( cos ) −<br />

2 ( − 0) 2<br />

The Lagrange equations of motion <strong>in</strong> the fulcrum frame are given by<br />

2 <br />

3


12.8. CORIOLIS FORCE 293<br />

Λ =0<br />

¨ − ˙ 2 − ( + )cos + ( − 0)=0<br />

Λ =0<br />

¨ + 2 ˙ ˙ ( + )<br />

+ s<strong>in</strong> =0<br />

<br />

These are identical to the Lagrange equations of motion derived <strong>in</strong> the <strong>in</strong>ertial frame.<br />

The can be used to derive the momenta <strong>in</strong> the non-<strong>in</strong>ertial fulcrum frame<br />

˜ = <br />

= ˙<br />

˙<br />

˜ = <br />

˙<br />

= 2 ˙<br />

which comprise only a part of the momenta derived <strong>in</strong> the <strong>in</strong>ertial frame. These partial fulcrum momenta<br />

lead to a Hamiltonian for the fulcum-frame of<br />

=˜ ˙ +˜ ˙ − = ˜2 <br />

2 + ˜ <br />

2 2 + 1 2 ( − 0) 2 − ( + ) cos <br />

Both and are time <strong>in</strong>dependent and thus the fulcrum Hamiltonian is a constant<br />

of motion <strong>in</strong> the fulcrum frame. However, does not equal the total energy which is <strong>in</strong>creas<strong>in</strong>g with<br />

time due to the acceleration of the fulcrum frame relative to the <strong>in</strong>ertial frame. This example illustrates that<br />

use of non-<strong>in</strong>ertial frames can simplify solution of accelerat<strong>in</strong>g systems.<br />

12.2 Example: Surface of rotat<strong>in</strong>g liquid<br />

F<strong>in</strong>d the shape of the surface of liquid <strong>in</strong> a bucket<br />

that rotates with angular speed as shown <strong>in</strong> the adjacent<br />

figure. Assume that the liquid is at rest <strong>in</strong> the<br />

frame of the bucket. Therefore, <strong>in</strong> the coord<strong>in</strong>ate system<br />

rotat<strong>in</strong>g with the bucket of liquid, the centrifugal force is<br />

important whereas the Coriolis, translational, and transverse<br />

forces are zero. The external force<br />

F = F 0 − g<br />

F’<br />

2<br />

where F 0 is the pressure which is perpendicular to the<br />

surface. At equilibrium the acceleration of the surface is<br />

zero that is<br />

a 00 =0=F 0 + (g − ω × (ω × r 0 ))<br />

mg<br />

mg<br />

The effective gravitational force is<br />

g =(g − ω × (ω × r 0 ))<br />

which must be perpendicular to the surface of the liquid s<strong>in</strong>ce F 0 is perpendicular to the surface of a fluid,<br />

and the net force is zero. In cyl<strong>in</strong>drical coord<strong>in</strong>ates this can be written as<br />

g = −bz + 2 bρ<br />

From the figure it can be deduced that<br />

By <strong>in</strong>tegration<br />

tan = <br />

= 2<br />

<br />

= 2<br />

2 2 + constant


294 CHAPTER 12. NON-INERTIAL REFERENCE FRAMES<br />

This is the equation of a paraboloid and corresponds to a parabolic gravitational equipotential energy surface.<br />

Astrophysicists build large parabolic mirrors for telescopes by cont<strong>in</strong>uously sp<strong>in</strong>n<strong>in</strong>g a large vat of glass while<br />

it solidifies. This is much easier than gr<strong>in</strong>d<strong>in</strong>g a large cyl<strong>in</strong>drical block of glass <strong>in</strong>to a parabolic shape.<br />

12.3 Example: The pirouette<br />

An <strong>in</strong>terest<strong>in</strong>g application of the Coriolis force is the problem of a sp<strong>in</strong>n<strong>in</strong>g ice skater or ballet dancer.<br />

Her angular frequency <strong>in</strong>creases when she draws <strong>in</strong> her arms. The conventional explanation is that angular<br />

momentum is conserved <strong>in</strong> the absence of any external forces which is correct. Thus s<strong>in</strong>ce her moment of<br />

<strong>in</strong>ertia decreases when she retracts her arms, her angular velocity must <strong>in</strong>crease to ma<strong>in</strong>ta<strong>in</strong> a constant<br />

angular momentum L = ω But this explanation does not address the question as to what are the forces<br />

that cause the angular frequency to <strong>in</strong>crease? The real radial forces the skater feels when she retracts her<br />

arms cannot directly lead to angular acceleration s<strong>in</strong>ce radial forces are perpendicular to the rotation. The<br />

follow<strong>in</strong>g derivation shows that the Coriolis force −2ω × v 00 acts tangentially to the radial retraction<br />

velocity of her arms lead<strong>in</strong>g to the angular acceleration required to ma<strong>in</strong>ta<strong>in</strong> constant angular momentum.<br />

Consider that a mass is mov<strong>in</strong>g radially at a velocity ˙ 00 then the Coriolis force <strong>in</strong> the rotat<strong>in</strong>g frame<br />

is<br />

F = −2ω × ṙ 00<br />

<br />

This Coriolis force leads to an angular acceleration of the mass of<br />

˙ω = −<br />

2ω × ṙ00 <br />

”<br />

()<br />

that is, the rotational frequency decreases if the radius is <strong>in</strong>creased. Note that, as shown <strong>in</strong> equation 1217<br />

˙ = ˙ 00 . This nonzero value of ˙ obviously leads to an azimuthal force <strong>in</strong> addition to the Coriolis force.<br />

Consider the rate of change of angular momentum for the rotat<strong>in</strong>g mass assum<strong>in</strong>g that the angular<br />

momentum comes purely from the rotation Then <strong>in</strong> the rotat<strong>in</strong>g frame<br />

ṗ 00 = (”2 ω)=2 00 ˙ 00 ω + 002 ˙ω<br />

Substitut<strong>in</strong>g equation for ˙ <strong>in</strong> the second term gives<br />

ṗ ” =2 00 ˙ 00 ω−2 00 ˙ 00 ω =0<br />

That is, the two terms cancel. Thus the angular momentum is conserved for this case where the velocity is<br />

radial. Note that, s<strong>in</strong>ce ” is assumed to be col<strong>in</strong>ear with then it is the same <strong>in</strong> both the stationary and<br />

rotat<strong>in</strong>g frames of reference and thus angular momentum is conserved <strong>in</strong> both frames. In addition, <strong>in</strong> the<br />

fixed frame, the angular momentum is conserved if no external torques are act<strong>in</strong>g as assumed above.<br />

Note that the rotational energy is<br />

Also the angular momentum is conserved, that is<br />

= 1 2 2<br />

p = ω = ˆω<br />

Substitut<strong>in</strong>g ω = p <br />

<br />

<strong>in</strong> the rotational energy gives<br />

= 2 <br />

2 = 2<br />

2<br />

Therefore the rotational energy actually <strong>in</strong>creases as the moment of <strong>in</strong>ertia decreases when the ice skater<br />

pulls her arms close to her body. This <strong>in</strong>crease <strong>in</strong> rotational energy is provided by the work done as the<br />

dancer pulls her arms <strong>in</strong>ward aga<strong>in</strong>st the centrifugal force.


12.9. ROUTHIAN REDUCTION FOR ROTATING SYSTEMS 295<br />

12.9 Routhian reduction for rotat<strong>in</strong>g systems<br />

The Routhian reduction technique, that was <strong>in</strong>troduced <strong>in</strong> chapter 86 is a hybrid variational approach. It<br />

was devised by Routh to handle the cyclic and non-cyclic variables separately <strong>in</strong> order to simultaneously<br />

exploit the differ<strong>in</strong>g advantages of the Hamiltonian and Lagrangian formulations. The Routhian reduction<br />

technique is a powerful method for handl<strong>in</strong>g rotat<strong>in</strong>g systems rang<strong>in</strong>g from galaxies to molecules, or deformed<br />

nuclei, as well as rotat<strong>in</strong>g mach<strong>in</strong>ery <strong>in</strong> eng<strong>in</strong>eer<strong>in</strong>g. A valuable feature of the Hamiltonian formulation is<br />

that it allows elim<strong>in</strong>ation of cyclic variables which reduces the number of degrees of freedom to be handled.<br />

As a consequence, cyclic variables are called ignorable variables <strong>in</strong> Hamiltonian mechanics. The Lagrangian,<br />

the Hamiltonian and the Routhian all are scalars under rotation and thus are <strong>in</strong>variant to rotation of the<br />

frame of reference. Note that often there are only two cyclic variables for a rotat<strong>in</strong>g system, that is, ˙θ = ω<br />

and the correspond<strong>in</strong>g canonical total angular momentum p = J.<br />

As mentioned <strong>in</strong> chapter 86, there are two possible Routhians that are useful for handl<strong>in</strong>g rotation frames<br />

of reference. For rotat<strong>in</strong>g systems the cyclic Routhian simplifies to<br />

( 1 ; ˙ 1 ˙ ; +1 ; ) = − = ω · J − (12.43)<br />

This Routhian behaves like a Hamiltonian for the ignorable cyclic coord<strong>in</strong>ates ω J Simultaneously it behaves<br />

like a negative Lagrangian for all the other coord<strong>in</strong>ates<br />

The non-cyclic Routhian complements <strong>in</strong> that it is def<strong>in</strong>ed as<br />

( 1 ; 1 ; ˙ +1 ˙ ; ) = − = − ω · J (12.44)<br />

This non-cyclic Routhian behaves like a Hamiltonian for all the non-cyclic variables and behaves like a<br />

negative Lagrangian for the two cyclic variables . S<strong>in</strong>ce the cyclic variables are constants of motion,<br />

then is a constant of motion that equals the energy <strong>in</strong> the rotat<strong>in</strong>g frame if is a constant of<br />

motion. However, does not equal the total energy s<strong>in</strong>ce the coord<strong>in</strong>ate transformation is time<br />

dependent, that is, the Routhian corresponds to the energy of the non-cyclic parts of the motion.<br />

For example, the Routhian for a system that is be<strong>in</strong>g cranked about the axis at some fixed<br />

angular frequency ˙ = with correspond<strong>in</strong>g total angular momentum p = J can be written as 1<br />

= − ω · J (12.45)<br />

= 1 2 hV · V + v” · v”+2V · v”+2V · (ω × r 0 )+2v” · (ω × r 0 )+(ω × r 0 ) 2i − ω · J + ()<br />

Note that is a constant of motion if <br />

<br />

=0 which is the case when the system is be<strong>in</strong>g cranked<br />

at a constant angular frequency. However the Hamiltonian <strong>in</strong> the rotat<strong>in</strong>g frame = − ω · J is given<br />

by = 6= s<strong>in</strong>ce the coord<strong>in</strong>ate transformation is time dependent. The canonical Hamilton<br />

equations for the fourth and fifth terms <strong>in</strong> the bracket can be identified with the Coriolis force 2ω × v 00 <br />

while the last term <strong>in</strong> the bracket is identified with the centrifugal force. That is, def<strong>in</strong>e<br />

≡− 1 2 (ω × r0 ) 2 (12.46)<br />

where the gradient of <br />

gives the usual centrifugal force.<br />

F = −∇ = 2 ∇ h 2 02 − (ω · r 0 ) 2i = £ 2 r 0 − (ω · r 0 ) ω ¤ = −ω × (ω × r 0 ) (12.47)<br />

The Routhian reduction method is used extensively <strong>in</strong> science and eng<strong>in</strong>eer<strong>in</strong>g to describe rotational<br />

motion of rigid bodies, molecules, deformed nuclei, and astrophysical objects. The cyclic variables describe<br />

the rotation of the frame and thus the Routhian = corresponds to the Hamiltonian for the<br />

non-cyclic variables <strong>in</strong> the rotat<strong>in</strong>g frame.<br />

1 For clarity sections 121 to 128 of this chapter adopted a nam<strong>in</strong>g convention that uses unprimed coord<strong>in</strong>ates with the<br />

subscript for the <strong>in</strong>ertial frame of reference, primed coord<strong>in</strong>ates with the subscript for the translat<strong>in</strong>g coord<strong>in</strong>ates, and<br />

double-primed coord<strong>in</strong>ates with the subscript for the translat<strong>in</strong>g plus rotat<strong>in</strong>g frame. For brevity the subsequent discussion<br />

omits the redundant subscripts s<strong>in</strong>ce the s<strong>in</strong>gle and double prime superscripts completely def<strong>in</strong>e the mov<strong>in</strong>g and<br />

rotat<strong>in</strong>g frames of reference.


296 CHAPTER 12. NON-INERTIAL REFERENCE FRAMES<br />

12.4 Example: Cranked plane pendulum<br />

The cranked plane pendulum, which is also called the rotat<strong>in</strong>g plane<br />

pendulum, comprises a plane pendulum that is cranked around a vertical<br />

axis at a constant angular velocity ˙ = as determ<strong>in</strong>ed by some<br />

external drive mechanism. The parameters are illustrated <strong>in</strong> the adjacent<br />

figure. The cranked pendulum nicely illustrates the advantages of<br />

work<strong>in</strong>g <strong>in</strong> a non-<strong>in</strong>ertial rotat<strong>in</strong>g frame for a driven rotat<strong>in</strong>g system.<br />

Although the cranked plane pendulum looks similar to the spherical pendulum,<br />

there is one very important difference; for the spherical pendulum<br />

= 2 s<strong>in</strong> 2 ˙ is a constant of motion and thus the angular velocity<br />

<br />

2 s<strong>in</strong> 2 <br />

varies with , i.e. ˙ = whereas for the cranked plane pendulum,<br />

the constant of motion is ˙ = and thus the angular momentum varies<br />

with i.e. = s<strong>in</strong> 2 . For the cranked plane pendulum, the energy<br />

must flow <strong>in</strong>to and out of the crank<strong>in</strong>g drive system that is provid<strong>in</strong>g the<br />

constra<strong>in</strong>t force to satisfy the equation of constra<strong>in</strong>t<br />

= ˙ − =0<br />

g<br />

Cranked plane pendulum that is<br />

cranked around the vertical axis<br />

with angular velocity ˙ = .<br />

The easiest way to solve the equations of motion for the cranked plane pendulum is to use generalized coord<strong>in</strong>ates<br />

to absorb the equation of constra<strong>in</strong>t and applied constra<strong>in</strong>t torque. This is done by <strong>in</strong>corporat<strong>in</strong>g the<br />

˙ = constra<strong>in</strong>t explicitly <strong>in</strong> the Lagrangian or Hamiltonian and solv<strong>in</strong>g for just <strong>in</strong> the rotat<strong>in</strong>g frame.<br />

Assum<strong>in</strong>g that ˙ = and us<strong>in</strong>g generalized coord<strong>in</strong>ates to absorb the crank<strong>in</strong>g constra<strong>in</strong>t forces, then<br />

the Lagrangian for the cranked pendulum can be written as.<br />

= 1 2 2 (˙ 2 +s<strong>in</strong> 2 2 )+ cos <br />

The momentum conjugate to is<br />

= <br />

˙ = 2 ˙<br />

Consider the Routhian = ˙ − = − ˙ which acts as a Hamiltonian <strong>in</strong> the rotat<strong>in</strong>g<br />

frame<br />

<br />

= ˙ − = − ˙ 2 =<br />

<br />

2 2 − 1 2 2 2 s<strong>in</strong> 2 − cos <br />

Note that if ˙ = is constant, then is a constant of motion for rotation about the axis s<strong>in</strong>ce<br />

it is <strong>in</strong>dependent of Also <br />

<br />

= − <br />

<br />

=0thus the energy <strong>in</strong> the rotat<strong>in</strong>g non-<strong>in</strong>ertial frame of the<br />

pendulum = = − ˙ is a constant of motion, but it does not equal the total energy s<strong>in</strong>ce<br />

the rotat<strong>in</strong>g coord<strong>in</strong>ate transformation is time dependent. The driver that cranks the system at a constant <br />

provides or absorbs the energy = = as changes <strong>in</strong> order to ma<strong>in</strong>ta<strong>in</strong> a constant .<br />

The Routhian can be used to derive the equations of motion us<strong>in</strong>g Hamiltonian mechanics.<br />

˙ = <br />

<br />

˙ = − <br />

<br />

= <br />

2<br />

= − s<strong>in</strong> <br />

∙1 − cos 2¸<br />

<br />

S<strong>in</strong>ce ˙ = 2¨ then the equation of motion is<br />

∙<br />

<br />

¨ +<br />

s<strong>in</strong> 1 − 2¸ cos =0 ()<br />

Assum<strong>in</strong>g that s<strong>in</strong> ≈ , then equation leads to l<strong>in</strong>ear harmonic oscillator solutions i about a m<strong>in</strong>imum<br />

at =0if the term <strong>in</strong> brackets is positive. That is, when the bracket<br />

h1 − cos 2 0 then equation <br />

corresponds to a harmonic oscillator with angular velocity Ω given by<br />

Ω 2 = ∙1 − cos 2¸<br />

<br />

m


12.9. ROUTHIAN REDUCTION FOR ROTATING SYSTEMS 297<br />

The adjacent figure shows the phase-space diagrams for a plane<br />

P<br />

pendulum rotat<strong>in</strong>g about a vertical axis at angular velocity for (a)<br />

p <br />

and (b) p <br />

<br />

The upper phase plot shows small when<br />

the square bracket of equation is positive and the the phase space<br />

trajectories are ellipses around the stable equilibrium po<strong>in</strong>t (0 0).<br />

As <strong>in</strong>creases the bracket becomes smaller and changes sign when<br />

2 cos = <br />

.Forlarger the bracket is negative lead<strong>in</strong>g to hyperbolic<br />

phase space trajectories around the ( )=(0 0) equilibrium<br />

po<strong>in</strong>t, that is, an unstable equilibrium po<strong>in</strong>t. However, new stable<br />

equilibrium po<strong>in</strong>ts now occur at angles ( )=(± 0 0) where<br />

(a)<br />

P<br />

cos 0 =<br />

<br />

<br />

That is, the equilibrium po<strong>in</strong>t (0 0) undergoes bifurcation<br />

as illustrated <strong>in</strong> the lower figure. These new equilibrium po<strong>in</strong>ts<br />

2<br />

arestableasillustratedbytheellipticaltrajectoriesaroundthese<br />

po<strong>in</strong>ts. It is <strong>in</strong>terest<strong>in</strong>g that these new equilibrium po<strong>in</strong>ts ± 0 move<br />

to larger angles given by 0 =<br />

<br />

<br />

beyond the bifurcation po<strong>in</strong>t<br />

<br />

2<br />

at<br />

<br />

=1. For low energy the mass oscillates about the m<strong>in</strong>imum<br />

2<br />

at = 0 whereas the motion becomes more complicated for higher<br />

(b)<br />

energy. The bifurcation corresponds to symmetry break<strong>in</strong>g s<strong>in</strong>ce,<br />

under spatial reflection, the equilibrium po<strong>in</strong>t is unchanged at low Phase-space diagrams for the plane<br />

rotational frequencies but it transforms from + 0 to − 0 once the pendulum cranked at angular velocity<br />

solution bifurcates, that is, the symmetry is broken. Also chaos can about a vertical axis. Figure () is<br />

occur at the separatrix that separates the bifurcation. Note that for <br />

while () is for .<br />

either the Lagrange multiplier approach, or the generalized force approach,<br />

can be used to determ<strong>in</strong>e the applied torque required to ensure a constant for the cranked pendulum.<br />

12.5 Example: Nucleon orbits <strong>in</strong> deformed nuclei<br />

Consider the rotation of axially-symmetric,<br />

prolate-deformed nucleus. Many nuclei have a prolate<br />

spheroidal shape, (the shape of a rugby ball)<br />

and they rotate perpendicular to the symmetry axis.<br />

In the non-<strong>in</strong>ertial body-fixed frame, pairs of nucleons,<br />

each with angular momentum are bound <strong>in</strong><br />

orbits with the projection of the angular momentum<br />

along the symmetry axis be<strong>in</strong>g conserved with value<br />

Ω = which is a cyclic variable. S<strong>in</strong>ce the nucleus<br />

is of dimensions 10 −14 quantization is important<br />

and the quantized b<strong>in</strong>d<strong>in</strong>g energies of the <strong>in</strong>dividual<br />

nucleons are separated by spac<strong>in</strong>gs ≤ 500<br />

The Lagrangian and Hamiltonian are scalars<br />

and can be evaluated <strong>in</strong> any coord<strong>in</strong>ate frame of<br />

reference. It is most useful to calculate the Hamiltonian<br />

for a deformed body <strong>in</strong> the non-<strong>in</strong>ertial rotat<strong>in</strong>g<br />

body-fixed frame of reference. The bodyfixed<br />

Hamiltonian corresponds to the Routhian<br />

<br />

= − ω · J<br />

Schematic diagram for the strong coupl<strong>in</strong>g of a<br />

nucleon to the deformation axis. The projection of <br />

on the symmetry axis is , andtheprojectionof is<br />

Ω. For axial symmetry Noether’s theroem gives that<br />

the projection of the angular momentum on the<br />

symmetry axis is a conserved quantity.<br />

where it is assumed that the deformed nucleus has the symmetry axis along the direction and rotates about<br />

the axis. S<strong>in</strong>ce the Routhian is for a non-<strong>in</strong>ertial rotat<strong>in</strong>g frame of reference it does not <strong>in</strong>clude the total<br />

energy but, if the shape is constant <strong>in</strong> time, then and the correspond<strong>in</strong>g body-fixed Hamiltonian<br />

are conserved and the energy levels for the nucleons bound <strong>in</strong> the spheroidal potential well can be calculated<br />

us<strong>in</strong>g a conventional quantum mechanical model.<br />

For a prolate spheroidal deformed potential well, the nucleon orbits that have the angular momentum<br />

nearly aligned to the symmetry axis correspond to nucleon trajectories that are restricted to the narrowest


298 CHAPTER 12. NON-INERTIAL REFERENCE FRAMES<br />

part of the spheroid, whereas trajectories with the angular momentum vector close to perpendicular to the<br />

symmetry axis have trajectories that probe the largest radii of the spheroid. The Heisenberg Uncerta<strong>in</strong>ty<br />

Pr<strong>in</strong>ciple, mentioned <strong>in</strong> chapter 3113, describes how orbits restricted to the smallest dimension will have<br />

the highest l<strong>in</strong>ear momentum, and correspond<strong>in</strong>g k<strong>in</strong>etic energy, and vise versa for the larger sized orbits.<br />

Thus the b<strong>in</strong>d<strong>in</strong>g energy of different nucleon trajectories <strong>in</strong> the spheroidal potential well depends on the angle<br />

between the angular momentum vector and the symmetry axis of the spheroid as well as the deformation of<br />

the spheroid. A quantal nuclear model Hamiltonian is solved for assumed spheroidal-shaped potential wells.<br />

The correspond<strong>in</strong>g orbits each have angular momenta j for which the projection of the angular momentum<br />

along the symmetry axis Ω is conserved, but the projection of j <strong>in</strong> the laboratory frame is not conserved<br />

s<strong>in</strong>ce the potential well is not spherically symmetric. However, the total Hamiltonian is spherically symmetric<br />

<strong>in</strong> the laboratory frame, which is satisfied by allow<strong>in</strong>g the deformed spheroidal potential well to rotate freely <strong>in</strong><br />

the laboratory frame, and then 2 and Ω all are conserved quantities. The attractive residual nucleonnucleon<br />

pair<strong>in</strong>g <strong>in</strong>teraction results <strong>in</strong> pairs of nucleons be<strong>in</strong>g bound <strong>in</strong> time-reversed orbits ( × ) 0 ,that<br />

is, with resultant total sp<strong>in</strong> zero, <strong>in</strong> this spheroidal nuclear potential. Excitation of an even-even nucleus<br />

can break one pair and then the total projection of the angular momentum along the symmetry axis is<br />

= |Ω 1 ± Ω 2 |, depend<strong>in</strong>g on whether the projections are parallel or antiparallel. More excitation energy<br />

can break several pairs and the projections cont<strong>in</strong>ue to be additive. The b<strong>in</strong>d<strong>in</strong>g energies calculated <strong>in</strong> the<br />

spheroidal potential well must be added to the rotational energy = J 2 2 to get the total energy, where<br />

J is the moment of <strong>in</strong>ertia. Nuclear structure measurements are <strong>in</strong> good agreement with the predictions of<br />

nuclear structure calculations that employ the Routhian approach.<br />

12.10 Effective gravitational force near the surface of the Earth<br />

Consider that the translational acceleration of the center of<br />

the Earth can be neglected, and thus a set of non-rotat<strong>in</strong>g<br />

axes through the center of the Earth can be assumed to be<br />

approximately an <strong>in</strong>ertial frame. The effects of the motion of<br />

the Earth around the Sun, or the motion of the Solar system<br />

<strong>in</strong> our Galaxy, are small compared with the effects due to the<br />

rotation of the Earth.<br />

Consider a rotat<strong>in</strong>g frame attached to the surface of the<br />

earth as shown <strong>in</strong> figure 125. The vector with respect to the<br />

center of the Earth r can be decomposed <strong>in</strong>to a vector to the<br />

orig<strong>in</strong> of the reference frame fixed to the surface of the Earth<br />

R plus the vector with respect to this surface reference frame<br />

r 0 <br />

r = R + r 0 (12.48)<br />

If the external force is separated <strong>in</strong>to the gravitational<br />

term g plus some other physical force F then the acceleration<br />

<strong>in</strong> the non-<strong>in</strong>ertial surface frame of reference is<br />

a 0 = F +g−(A +2ω × v0 + ω × (ω × r 0 )+ ˙ω × r 0 ) (12.49)<br />

O’<br />

r’<br />

x 3<br />

x 2<br />

r<br />

O x 1<br />

Figure 12.5: Rotat<strong>in</strong>g frame at the surface of<br />

the Earth.<br />

P<br />

But<br />

V =<br />

µ R<br />

=<br />

<br />

<br />

s<strong>in</strong>ce <strong>in</strong> the rotat<strong>in</strong>g frame ¡ ¢<br />

R<br />

<br />

=0 Also the acceleration<br />

<br />

µ V<br />

A =<br />

<br />

<br />

µ V<br />

=<br />

<br />

µ R<br />

+ ω × R = ω × R (12.50)<br />

<br />

<br />

<br />

+ ω × V = ω × (ω × R) (12.51)


12.10. EFFECTIVE GRAVITATIONAL FORCE NEAR THE SURFACE OF THE EARTH 299<br />

s<strong>in</strong>ce ¡ ¢<br />

V<br />

<br />

=0 Substitut<strong>in</strong>g this <strong>in</strong>to the above equation gives<br />

<br />

a 0 = F + g − (2ω × v0 + ω × (ω × [r 0 + R]) + ˙ω × r 0 )<br />

= F + g − (2ω × v0 + ω × (ω × r)+ ˙ω × r 0 )<br />

where r is with respect to the center of the Earth. This is as expected directly from equation 1236. S<strong>in</strong>ce<br />

the angular frequency of the earth is a constant then ˙ × r 0 =0 Thus the acceleration can be written as<br />

a 0 = F +[g − ω × (ω × r)] − 2ω × v0 (12.52)<br />

The term <strong>in</strong> the square brackets comb<strong>in</strong>es the gravitational acceleration plus the centrifugal acceleration.<br />

A measurement of the Earth’s gravitational acceleration<br />

actually measures the term <strong>in</strong> the square brackets<br />

<strong>in</strong> equation 1252, thatis,aneffective gravitational<br />

acceleration where<br />

g = g − ω × (ω × r) (12.53)<br />

near the surface of the earth r ≈ R. Theeffective gravitational<br />

force does not po<strong>in</strong>t towards the center of the<br />

Earth as shown <strong>in</strong> figure 126. A plumb l<strong>in</strong>e po<strong>in</strong>ts,<br />

or an object falls, <strong>in</strong> the direction of g The shape<br />

of the earth is such that the Earth’s surface is perpendicular<br />

to g . Thisisthereasonwhytheearthis<br />

distorted <strong>in</strong>to an oblate ellipsoid, that is, it is flattened<br />

at the poles.<br />

The angle between g and the l<strong>in</strong>e po<strong>in</strong>t<strong>in</strong>g<br />

to the center of the earth is dependent on the latitude<br />

= 2<br />

− Note that the colatitude is taken to be zero<br />

at the North pole whereas the latitude is taken to<br />

be zero at the equator. The angle can be estimated<br />

by assum<strong>in</strong>g that 0 then the centrifugal term<br />

then can be approximated by<br />

g<br />

Figure 12.6: Effective gravitational acceleration.<br />

g<br />

|ω × (ω × r)| ≈ 2 s<strong>in</strong> = 2 cos (12.54)<br />

This is quite small for the Earth s<strong>in</strong>ce =073 × 10 −4 and = 6371 lead<strong>in</strong>g to a correction<br />

term 2 cos =003 cos 2 S<strong>in</strong>ce<br />

<br />

= 2 cos s<strong>in</strong> (12.55)<br />

and<br />

= − 2 cos 2 (12.56)<br />

Then the angle between g and g is given by<br />

' tan = <br />

<br />

<br />

<br />

= 2 cos s<strong>in</strong> <br />

− 2 cos 2 <br />

(12.57)<br />

This has a maximum value at =45 which is =00088 ◦ .


300 CHAPTER 12. NON-INERTIAL REFERENCE FRAMES<br />

12.11 Free motion on the earth<br />

The calculation of trajectories for objects as they move near<br />

the surface of the earth is frequently required for many applications.<br />

Such calculations require <strong>in</strong>clusion of the non<strong>in</strong>ertial<br />

Coriolis force. In the frame of reference fixed to<br />

the earth’s surface, assum<strong>in</strong>g that air resistance and other<br />

forces can be neglected, then the acceleration equals<br />

y<br />

(North)<br />

z (vertical)<br />

a 0 = g − 2ω × v 0 (12.58)<br />

Neglect the centrifugal correction term s<strong>in</strong>ce it is very small,<br />

that is, let g = g. Us<strong>in</strong>gthecoord<strong>in</strong>ateaxisshown<strong>in</strong><br />

figure 127, the surface-frame vectors have components<br />

x (East)<br />

ω =0 b i 0 + cos b j 0 + s<strong>in</strong> b k 0 (12.59)<br />

Equator<br />

and<br />

g = −k b0 (12.60)<br />

Thus the Coriolis term is<br />

bi 0 j b0 k b0<br />

2ω × v 0 = 2<br />

0 cos s<strong>in</strong> <br />

¯ <br />

0 0 <br />

¯¯¯¯¯¯ 0<br />

h³<br />

= 2 0 cos − 0 s<strong>in</strong> ´<br />

bi0 +<br />

Therefore the equations of motion are<br />

³<br />

³<br />

i<br />

0 s<strong>in</strong> ´<br />

bj0 − 0 cos ´<br />

bk<br />

0<br />

Figure 12.7: Rotat<strong>in</strong>g frame fixedonthesurface<br />

of the Earth.<br />

¨r 0 = − b k 0 −2[ b i 0 ( ˙ 0 cos − ˙ 0 s<strong>in</strong> )+ b j 0 ˙ 0 s<strong>in</strong> − b k 0 ˙ 0 cos ] (12.61)<br />

That is, the components of this equation of motion are<br />

Integrat<strong>in</strong>g these differential equations gives<br />

¨ 0 = −2 ( ˙ 0 cos − ˙ 0 s<strong>in</strong> ) (12.62)<br />

¨ 0 = −2 ˙ 0 s<strong>in</strong> <br />

¨ 0 = − +2 ˙ 0 cos <br />

˙ 0 = −2 ( 0 cos − 0 s<strong>in</strong> )+ ˙ 0 0 (12.63)<br />

˙ 0 = −2 0 s<strong>in</strong> + ˙ 0<br />

0<br />

˙ 0 = − +2 0 cos + ˙ 0 0<br />

where ˙ 0 0 ˙ 0 0 ˙ 0 0 are the <strong>in</strong>itial velocities. Substitut<strong>in</strong>g the above velocity relations <strong>in</strong>to the equation of motion<br />

for ¨ gives<br />

¨ 0 =2 cos − 2 ( ˙ 0 0 cos − ˙ 0 0 s<strong>in</strong> ) − 4 2 0 (12.64)<br />

The last term 4 2 is small and can be neglected lead<strong>in</strong>g to a simple uncoupled second-order differential<br />

equation <strong>in</strong> . Integrat<strong>in</strong>g this twice assum<strong>in</strong>g that 0 0 = 0 0 = 0 0 =0 plus the fact that 2 cos and<br />

2 ( ˙ 0 0 cos − ˙ 0 0 s<strong>in</strong> ) are constant, gives<br />

0 = 1 3 3 cos − 2 ( ˙ 0 0 cos − ˙ 0 0 s<strong>in</strong> )+ ˙ 0 0 (12.65)<br />

Similarly,<br />

0 = ¡ ˙ 0 0 − ˙ 0 0 2 s<strong>in</strong> ¢ (12.66)<br />

0 = − 1 2 2 + ˙ 0 0 + ˙ 0 0 2 cos (12.67)


12.11. FREE MOTION ON THE EARTH 301<br />

Consider the follow<strong>in</strong>g special cases;<br />

12.6 Example: Free fall from rest<br />

Assume that an object falls a height start<strong>in</strong>g from rest at =0, =0, =0, = . Then<br />

Substitut<strong>in</strong>g for gives<br />

0 = 1 3 3 cos <br />

0 = 0<br />

0 = − 1 2 2<br />

0 = 1 3 cos s<br />

8 3<br />

Thus the object drifts eastward as a consequence of the earth’s rotation. Note that relative to the fixed frame<br />

it is obvious that the angular velocity of the body must <strong>in</strong>crease as it falls to compensate for the reduced<br />

distance from the axis of rotation <strong>in</strong> order to ensure that the angular momentum is conserved.<br />

12.7 Example: Projectile fired vertically upwards<br />

An upward fired projectile with <strong>in</strong>itial velocities ˙ 0 0 = ˙ 0 0 =0and ˙ 0 0 = 0 leads to the relations<br />

0 = 1 3 3 cos − 2 0 cos <br />

0 =0<br />

<br />

0 = − 1 2 2 + 0 <br />

Solv<strong>in</strong>g for when 0 =0gives =0 and = 20<br />

<br />

Also s<strong>in</strong>ce the maximum height that the projectile<br />

reachesisrelatedby<br />

0 = p 2<br />

then the f<strong>in</strong>al deflection is<br />

s<br />

0 = − 4 3 cos 8 3<br />

<br />

Thus the body drifts westwards.<br />

12.8 Example: Motion parallel to Earth’s surface<br />

For motion <strong>in</strong> the horizontal 0 − 0 plane the deflection is always to the right <strong>in</strong> the northern hemisphere<br />

of the Earth s<strong>in</strong>ce the vertical component of is upwards and thus −2 −→ ω × −→ v 0 po<strong>in</strong>ts to the right. In the<br />

southern hemisphere the vertical component of is downward and thus −2 −→ ω × −→ v 0 po<strong>in</strong>ts to the left. This<br />

is also shown us<strong>in</strong>g the above relations for the case of a projectile fired upwards <strong>in</strong> an easterly direction with<br />

components 0 0 0 0 0 The resultant displacements are<br />

0 = 1 3 3 cos − 2 ˙ 0 0 cos + ˙ 0 0<br />

Similarly,<br />

0 = − ˙ 0 0 2 s<strong>in</strong> <br />

0 = − 1 2 2 + ˙ 0 0 + ˙ 0 0 2 cos <br />

The trajectory is non-planar and, <strong>in</strong> the northern hemisphere, the projectile drifts to the right, that is<br />

southerly.<br />

In the battle of the River de la Plata, dur<strong>in</strong>g World War 2, the gunners on the British light cruisers<br />

Exeter, Ajax and Achilles found that their accurately aimed salvos aga<strong>in</strong>st the German pocket battleship Graf<br />

Spee were fall<strong>in</strong>g 100 yards to the left. The designers of the gun sight<strong>in</strong>g mechanisms had corrected for the<br />

Coriolis effect assum<strong>in</strong>g the ships would fight at latitudes near 50 north, not 50 south.


302 CHAPTER 12. NON-INERTIAL REFERENCE FRAMES<br />

12.12 Weather systems<br />

Weather systems on Earth provide a classic example of motion <strong>in</strong> a rotat<strong>in</strong>g coord<strong>in</strong>ate system. In the<br />

northern hemisphere, air flow<strong>in</strong>g<strong>in</strong>toalow-pressureregionisdeflected to the right caus<strong>in</strong>g counterclockwise<br />

circulation, whereas air flow<strong>in</strong>g out of a high-pressure region is deflected to the right caus<strong>in</strong>g a clockwise<br />

circulation. Trade w<strong>in</strong>ds on the Earth result from air ris<strong>in</strong>g or s<strong>in</strong>k<strong>in</strong>g due to thermal activity comb<strong>in</strong>ed<br />

with the Coriolis effect. Similar behavior is observed on other planets such as the Red Spot on Jupiter.<br />

For a fluid or gas, equation (1236) can be written <strong>in</strong> terms of the fluid density <strong>in</strong> the form<br />

a” =−∇ − [2ω × v” − ω × (ω × r 0 )] (12.68)<br />

where the translational acceleration A, the gravitational force, and the azimuthal acceleration ( ˙ω × r 0 ) terms<br />

are ignored. The external force per unit volume equals the pressure gradient −∇ while ω is the rotation<br />

vector of the earth.<br />

In fluid flow, the Rossby number is def<strong>in</strong>ed to be<br />

<strong>in</strong>ertial force<br />

=<br />

Coriolis force ≈ a”<br />

(12.69)<br />

2ω × v”<br />

For large dimensional pressure systems <strong>in</strong> the atmosphere, e.g. ' 1000, the Rossby number is ∼ 01<br />

and thus the Coriolis force dom<strong>in</strong>ates and the radial acceleration can be neglected. This leads to a flow<br />

velocity ' 10 which is perpendicular to the pressure gradient ∇ ,thatis,theairflows horizontally<br />

parallel to the isobars of constant pressure which is called geostrophic flow. For much smaller dimension<br />

systems, such as at the wall of a hurricane, ' 50, and ' 50 the Rossby number ' 10 and<br />

the Coriolis effect plays a much less significant role compared to the balance between the radial centrifugal<br />

forces and the pressure gradient. The same situation of the Coriolis forces be<strong>in</strong>g <strong>in</strong>significant occurs for most<br />

small-scale vortices such as tornadoes, typical thermal vortices <strong>in</strong> the atmosphere, and for water dra<strong>in</strong><strong>in</strong>g a<br />

bath tub.<br />

12.12.1 Low-pressure systems:<br />

It is <strong>in</strong>terest<strong>in</strong>g to analyze the motion of air circulat<strong>in</strong>g<br />

around a low pressure region at large radii where<br />

the motion is tangential. As shown <strong>in</strong> figure 128,<br />

a parcel of air circulat<strong>in</strong>g anticlockwise around the<br />

low with velocity <strong>in</strong>volves a pressure difference ∆<br />

act<strong>in</strong>g on the surface area plus the centrifugal and<br />

Coriolis forces. Assum<strong>in</strong>g that these forces are balanced<br />

such that a” ' 0 then equation 1268 simplifies<br />

to<br />

2<br />

= 1 ∇ − 2 s<strong>in</strong> <br />

<br />

(12.70)<br />

where the latitude = −. Thustheforceequation<br />

can be written<br />

1 <br />

= 2<br />

+2 s<strong>in</strong> <br />

<br />

(12.71)<br />

It is apparent that the comb<strong>in</strong>ed outward Coriolis<br />

force plus outward centrifugal force, act<strong>in</strong>g on the<br />

Figure 12.8: Air flow and pressures around a lowpressure<br />

region.<br />

circulat<strong>in</strong>g air, can support a large pressure gradient.<br />

The tangential velocity can be obta<strong>in</strong>ed by solv<strong>in</strong>g this equation to give<br />

s<br />

= ( s<strong>in</strong> ) 2 + <br />

− s<strong>in</strong> <br />

<br />

(12.72)<br />

Note that the velocity equals zero when =0assum<strong>in</strong>g that <br />

<br />

maximum at a radius<br />

Low<br />

is f<strong>in</strong>ite. That is, the velocity reaches a<br />

= 1 4 (1 + 1 <br />

s<strong>in</strong> ) (12.73)<br />

S


12.12. WEATHER SYSTEMS 303<br />

Figure 12.9: Hurricane Katr<strong>in</strong>a over the Gulf of Mexico on 28 August 2005. [Published by the NOAA]<br />

which occurs at the wall of the eye of the circulat<strong>in</strong>g low-pressure system.<br />

Low pressure regions are produced by heat<strong>in</strong>g of air caus<strong>in</strong>g it to rise and result<strong>in</strong>g <strong>in</strong> an <strong>in</strong>flow of air<br />

to replace the ris<strong>in</strong>g air. Hurricanes form over warm water when the temperature exceeds 26 ◦ and the<br />

moisture levels are above average. They are created at latitudes between 10 ◦ −15 ◦ where the sea is warmest,<br />

but not closer to the equator where the Coriolis force drops to zero. About 90% of the heat<strong>in</strong>g of the air comes<br />

from the latent heat of vaporization due to the ris<strong>in</strong>g warm moist air condens<strong>in</strong>g <strong>in</strong>to water droplets <strong>in</strong> the<br />

cloud similar to what occurs <strong>in</strong> thunderstorms. For hurricanes <strong>in</strong> the northern hemisphere, the air circulates<br />

anticlockwise <strong>in</strong>wards. Near the wall of the eye of the hurricane, the air rises rapidly to high altitudes at<br />

which it then flows clockwise and outwards and subsequently back down <strong>in</strong> the outer reaches of the hurricane.<br />

Both the w<strong>in</strong>d velocity and pressure are low <strong>in</strong>side the eye which can be cloud free. The strongest w<strong>in</strong>ds<br />

are <strong>in</strong> vortex surround<strong>in</strong>g the eye of the hurricane, while weak w<strong>in</strong>ds exist <strong>in</strong> the counter-rotat<strong>in</strong>g vortex of<br />

s<strong>in</strong>k<strong>in</strong>g air that occurs far outside the hurricane.<br />

Figure 129 shows the satellite picture of the hurricane Katr<strong>in</strong>a, recorded on 28 August 2005. The eye of<br />

the hurricane is readily apparent <strong>in</strong> this picture. The central pressure was 90200 2 (902) compared<br />

with the standard atmospheric pressure of 101300 2 (1013). This 111 pressure difference produced<br />

steady w<strong>in</strong>ds <strong>in</strong> Katr<strong>in</strong>a of 280 ( 175) withgustsupto344 whichresulted<strong>in</strong>1833 fatalities.<br />

Tornadoes are another example of a vortex low-pressuresystemthataretheoppositeextreme<strong>in</strong>both<br />

size and duration compared with a hurricane. Tornadoes may last only ∼ 10 m<strong>in</strong>utes and be quite small <strong>in</strong><br />

radius. Pressure drops of up to 100 have been recorded, but s<strong>in</strong>ce they may only be a few 100 meters <strong>in</strong><br />

diameter, the pressure gradient can be much higher than for hurricanes lead<strong>in</strong>g to localized w<strong>in</strong>ds thought to<br />

approach 500. Unfortunately, the <strong>in</strong>strumentation and build<strong>in</strong>gs hit by a tornado often are destroyed<br />

mak<strong>in</strong>g study difficult. Note that the the pressure gradient <strong>in</strong> small diameter of rope tornadoes is much<br />

more destructive than for larger 14 mile diameter tornadoes, which results <strong>in</strong> stronger w<strong>in</strong>ds.


304 CHAPTER 12. NON-INERTIAL REFERENCE FRAMES<br />

12.12.2 High-pressure systems:<br />

In contrast to low-pressure systems, high-pressure systems are very different <strong>in</strong> that the Coriolis force po<strong>in</strong>ts<br />

<strong>in</strong>ward oppos<strong>in</strong>g the outward pressure gradient and centrifugal force. That is,<br />

2<br />

=2 s<strong>in</strong> − 1 <br />

(12.74)<br />

<br />

which gives that<br />

s<br />

= s<strong>in</strong> − ( s<strong>in</strong> ) 2 − <br />

(12.75)<br />

<br />

This implies that the maximum pressure gradient plus centrifugal force supported by the Coriolis force is<br />

<br />

≤ ( s<strong>in</strong> )2 (12.76)<br />

As a consequence, high pressure regions tend to have weak pressure gradients and light w<strong>in</strong>ds <strong>in</strong> contrast to<br />

the large pressure gradients plus concomitant damag<strong>in</strong>g w<strong>in</strong>ds possible for low pressure systems.<br />

The circulation behavior, exhibited by weather patterns, also applies to ocean currents and other liquid<br />

flow on earth. However, the residual angular momentum of the liquid often can overcome the Coriolis terms.<br />

Thus often it will be found experimentally that water exit<strong>in</strong>g the bathtub does not circulate anticlockwise <strong>in</strong><br />

the northern hemisphere as predicted by the Coriolis force. This is because it was not stationary orig<strong>in</strong>ally,<br />

but rotat<strong>in</strong>g slowly.<br />

Reliable prediction of weather is an extremely difficult, complicated and challeng<strong>in</strong>g task, which is of considerable<br />

importance <strong>in</strong> modern life. As discussed <strong>in</strong> chapter 168, fluid flow can be much more complicated<br />

thanassumed<strong>in</strong>thisdiscussionofairflow and weather. Both turbulent and lam<strong>in</strong>ar flow are possible. As a<br />

consequence, computer simulations of weather phenomena are difficult because the air flow can be turbulent<br />

and the transition from order to chaotic flow is very sensitive to the <strong>in</strong>itial conditions. Typically the air<br />

flow can <strong>in</strong>volve both macroscopic ordered coherent structures over a wide dynamic range of dimensions,<br />

coexist<strong>in</strong>g with chaotic regions. Computer simulations of fluid flow often are performed based on Lagrangian<br />

mechanics to exploit the scalar properties of the Lagrangian. Ordered coherent structures, rang<strong>in</strong>g from<br />

microscopic bubbles to hurricanes, can be recognized by exploit<strong>in</strong>g Lyapunov exponents to identify the ordered<br />

motion buried <strong>in</strong> the underly<strong>in</strong>g chaos. Thus the techniques discussed <strong>in</strong> classical mechanics are of<br />

considerable importance outside of physics.<br />

12.13 Foucault pendulum<br />

A classic example of motion <strong>in</strong> non-<strong>in</strong>ertial frames is the rotation of the Foucault pendulum on the surface of<br />

the earth. The Foucault pendulum is a spherical pendulum with a long suspension that oscillates <strong>in</strong> the −<br />

plane with sufficiently small amplitude that the vertical velocity ˙ is negligible. Assume that the pendulum<br />

is a simple pendulum of length and mass as shown <strong>in</strong> figure 1210. The equation of motion is given by<br />

¨r = g + T − 2Ω × ṙ (12.77)<br />

<br />

where <br />

is the acceleration produced by the tension <strong>in</strong> the pendulum suspension and the rotation vector<br />

of the earth is designated by Ω to avoid confusion with the oscillation frequency of the pendulum . The<br />

effective gravitational acceleration g is given by<br />

g = g 0 − Ω × [Ω × (r + R)] (12.78)<br />

that is, the true gravitational field g 0 corrected for the centrifugal force.<br />

Assume the small angle approximation for the pendulum deflection angle , then = cos ' and<br />

= , thus ' . Thenhasshown<strong>in</strong>figure 1210, the horizontal components of the restor<strong>in</strong>g force<br />

are<br />

= − (12.79)<br />

<br />

= − <br />

(12.80)


12.13. FOUCAULT PENDULUM 305<br />

S<strong>in</strong>ce g is vertical, and neglect<strong>in</strong>g terms <strong>in</strong>volv<strong>in</strong>g ˙ then evaluat<strong>in</strong>g<br />

the cross product <strong>in</strong> equation (1278) simplifies to<br />

z<br />

¨ = − <br />

¨ = − <br />

+2˙Ω cos (12.81)<br />

− 2 ˙Ω cos (12.82)<br />

where is the colatitude which is related to the latitude by<br />

cos =s<strong>in</strong> (12.83)<br />

The natural angular frequency of the simple pendulum is<br />

l<br />

y<br />

x<br />

l<br />

r <br />

0 =<br />

<br />

(12.84)<br />

x<br />

y<br />

l<br />

mg<br />

while the component of the earth’s angular velocity is<br />

Ω = Ω cos (12.85)<br />

Thus equations 1281 and 1282 can be written as<br />

Figure 12.10: Foucault pendulum.<br />

¨ − 2Ω ˙ + 2 0 = 0<br />

¨ − 2Ω ˙ + 2 0 = 0 (12.86)<br />

These are two coupled equations that can be solved by mak<strong>in</strong>g a coord<strong>in</strong>ate transformation.<br />

Def<strong>in</strong>e a new coord<strong>in</strong>ate that is a complex number<br />

= + (12.87)<br />

Multiply the second of the coupled equations 1286 by and add to the first equation gives<br />

which can be written as a differential equation for <br />

(¨ + ¨)+2Ω ( ˙ + ˙)+ 2 0 ( + ) =0<br />

¨ +2Ω ˙ + 2 0 =0 (12.88)<br />

Note that the complex number conta<strong>in</strong>s the same <strong>in</strong>formation regard<strong>in</strong>g the position <strong>in</strong> the − plane<br />

as equations 1286. The plot of <strong>in</strong> the complex plane, the Argand diagram, is a birds-eye view of the<br />

position coord<strong>in</strong>ates ( ) of the pendulum. This second-order homogeneous differential equation has two<br />

<strong>in</strong>dependent solutions that can be derived by guess<strong>in</strong>g a solution of the form<br />

Substitut<strong>in</strong>g equation 1289 <strong>in</strong>to 1288 gives that<br />

That is<br />

() = − (12.89)<br />

2 − 2Ω − 2 =0<br />

If the angular velocity of the pendulum 0 Ω, then<br />

q<br />

= Ω ± Ω 2 + 2 0 (12.90)<br />

Thus the solution is of the form<br />

This can be written as<br />

' Ω ± 0 (12.91)<br />

() = −Ω ( + 0 + − 0 ) (12.92)<br />

() = −Ω cos( 0 + ) (12.93)


306 CHAPTER 12. NON-INERTIAL REFERENCE FRAMES<br />

where the phase and amplitude depend on the <strong>in</strong>itial conditions. Thus the plane of oscillation of the<br />

pendulum is def<strong>in</strong>ed by the ratio of the and coord<strong>in</strong>ates, that is the phase angle Ω This phase angle<br />

rotates with angular velocity Ω where<br />

Ω = Ω cos = Ω s<strong>in</strong> (12.94)<br />

At the north pole the earth rotates under the pendulum with angular velocity Ω and the axis of the<br />

pendulum is fixed <strong>in</strong> an <strong>in</strong>ertial frame of reference. At lower latitudes, the pendulum precesses at the lower<br />

angular frequency Ω = Ω s<strong>in</strong> that goes to zero at the equator. For example, <strong>in</strong> Rochester, NY, =43 ◦ <br />

and therefore a Foucault pendulum precesses at Ω =0682Ω. That is, the pendulum precesses 2455 ◦ /day.<br />

12.14 Summary<br />

This chapter has focussed on describ<strong>in</strong>g motion <strong>in</strong> non-<strong>in</strong>ertial frames of reference. It has been shown that the<br />

force and acceleration <strong>in</strong> non-<strong>in</strong>ertial frames can be related us<strong>in</strong>g either Newtonian or Lagrangian mechanics<br />

by <strong>in</strong>troduc<strong>in</strong>g additional <strong>in</strong>ertial forces <strong>in</strong> the non-<strong>in</strong>ertial reference frame.<br />

Translational acceleration of a reference frame In a primed frame, that is undergo<strong>in</strong>g translational<br />

acceleration A the motion <strong>in</strong> this non-<strong>in</strong>ertial frame can be calculated by addition of an <strong>in</strong>ertial force -A,<br />

that leads to an equation of motion<br />

a 0 = F − A<br />

(126)<br />

Note that the primed frame is an <strong>in</strong>ertial frame if A =0.<br />

Rotat<strong>in</strong>g reference frame It was shown that the time derivatives of a general vector G <strong>in</strong> both an<br />

<strong>in</strong>ertial frame and a rotat<strong>in</strong>g reference frame are related by<br />

µ µ <br />

G G<br />

=<br />

<br />

<br />

+ ω × G (1216)<br />

<br />

<br />

where the ω × G term orig<strong>in</strong>ates from the fact that the unit vectors <strong>in</strong> the rotat<strong>in</strong>g reference frame are time<br />

dependent with respect to the <strong>in</strong>ertial frame.<br />

Reference frame undergo<strong>in</strong>g both rotation and translation Both Newtonian and Lagrangian mechanics<br />

were used to show that for the case of translational acceleration plus rotation, the effective force <strong>in</strong><br />

the non-<strong>in</strong>ertial (double-primed) frame can be written as<br />

F = a 00 = F − (A + ω × V +2ω × v 00 + ω × (ω × r 0 )+ ˙ω × r 0 ) (1228 1236)<br />

These <strong>in</strong>ertial correction forces result from describ<strong>in</strong>g the system us<strong>in</strong>g a non-<strong>in</strong>ertial frame. These <strong>in</strong>ertial<br />

forces are felt when <strong>in</strong> the rotat<strong>in</strong>g-translat<strong>in</strong>g frame of reference. Thus the notion of these <strong>in</strong>ertial forces<br />

can be very useful for solv<strong>in</strong>g problems <strong>in</strong> non-<strong>in</strong>ertial frames. For the case of rotat<strong>in</strong>g frames, two important<br />

<strong>in</strong>ertial forces are the centrifugal force, −ω × (ω × r 0 ) and the Coriolis force −2ω × v 00 .<br />

Routhian reduction for rotat<strong>in</strong>g systems It was shown that for non-<strong>in</strong>ertial systems, identical equations<br />

of motion are derived us<strong>in</strong>g Newtonian, Lagrangian, Hamiltonian, and Routhian mechanics.<br />

Terrestrial manifestations of rotation Examples of motion <strong>in</strong> rotat<strong>in</strong>g frames presented <strong>in</strong> the chapter<br />

<strong>in</strong>cluded projectile motion with respect to the surface of the Earth, rotation alignment of nucleons <strong>in</strong> rotat<strong>in</strong>g<br />

nuclei, and weather phenomena.


12.14. SUMMARY 307<br />

Problems<br />

1. Consider a fixed reference frame and a rotat<strong>in</strong>g frame 0 . The orig<strong>in</strong>s of the two coord<strong>in</strong>ate systems always<br />

co<strong>in</strong>cide. By carefully draw<strong>in</strong>g a diagram, derive an expression relat<strong>in</strong>g the coord<strong>in</strong>ates of a po<strong>in</strong>t <strong>in</strong> the two<br />

systems. (This was covered <strong>in</strong> Chapter 2, but it is worth review<strong>in</strong>g now.<br />

2. The effective force observed <strong>in</strong> a rotat<strong>in</strong>g coord<strong>in</strong>ate system is given by equation 1228.<br />

(a) What is the significance of each term <strong>in</strong> this expression?<br />

(b) Suppose you wanted to measure the gravitational force, both magnitude and direction, on a body of mass<br />

at rest on the surface of the Earth. What terms <strong>in</strong> the effective force can be neglected?<br />

(c) Suppose you wanted to calculate the deflection of a projectile fired horizontally along the Earth’s surface.<br />

What terms <strong>in</strong> the effective force can be neglected?<br />

(d) Suppose you wanted to calculate the effective force on a small block of mass placed on a frictionless<br />

turntable rotat<strong>in</strong>g with a time-dependent angular velocity (). Whatterms<strong>in</strong>theeffective force can be<br />

neglected?<br />

3. A plumb l<strong>in</strong>e is carried along <strong>in</strong> a mov<strong>in</strong>g tra<strong>in</strong>, with the mass of the plumb bob. Neglect any effects due to<br />

therotationoftheEarthandwork<strong>in</strong>thenon<strong>in</strong>ertial frame of reference of the tra<strong>in</strong>.<br />

(a) F<strong>in</strong>d the tension <strong>in</strong> the cord and the deflection from the local vertical if the tra<strong>in</strong> is mov<strong>in</strong>g with constant<br />

acceleration 0 .<br />

(b) F<strong>in</strong>d the tension <strong>in</strong> the cord and the deflection from the local vertical if the tra<strong>in</strong> is round<strong>in</strong>g a curve of<br />

radius with constant speed 0 .<br />

4. A bead on a rotat<strong>in</strong>g rod is free to slide without friction. The rod has a length and rotates about its end<br />

with angular velocity . The bead is <strong>in</strong>itially released from rest (relative to the rod) at the midpo<strong>in</strong>t of the<br />

rod.<br />

(a) F<strong>in</strong>d the displacement of the bead along the wire as a function of time.<br />

(b) F<strong>in</strong>d the time when the bead leaves the end of the rod.<br />

(c) F<strong>in</strong>d the velocity (relative to the rod) of the bead when it leaves the end of the rod.<br />

5. Here is a “thought experiment” for you to consider. Suppose you are <strong>in</strong> a small sailboat of mass at the<br />

Earth’s equator. At the equator there is very little w<strong>in</strong>d (this is known as the “equatorial doldrums”), so your<br />

sailboat is, more or less, sitt<strong>in</strong>g still. You have a small anchor of mass on deck and a s<strong>in</strong>gle mast of height<br />

<strong>in</strong> the middle of the boat. How can you use the anchor to put the boat <strong>in</strong>to motion? In which direction will<br />

the boat move?<br />

6. Does water really flow <strong>in</strong> the other direction when you flush a toilet <strong>in</strong> the southern hemisphere? What (if<br />

anyth<strong>in</strong>g) does the Coriolis force have to do with this?<br />

7. We are presently at a latitude (with respect to the equator) and Earth is rotat<strong>in</strong>g with constant angular<br />

velocity . Consider the follow<strong>in</strong>g two scenarios: Scenario A:Aparticleisthrownupwardwith<strong>in</strong>itialspeed<br />

0 . Scenario B: An identical particle is dropped (at rest) from the maximum height of the particle <strong>in</strong> Scenario<br />

A. Circle all the true statements regard<strong>in</strong>g the Coriolis deflection assum<strong>in</strong>g that the particles have landed for<br />

a) and b), .<br />

(a) The magnitude is greater <strong>in</strong> A than <strong>in</strong> B.<br />

(b) The direction <strong>in</strong> A and B are the same.<br />

(c) The direction <strong>in</strong> A does not change throughout flight.


308 CHAPTER 12. NON-INERTIAL REFERENCE FRAMES<br />

8. If a projectile is fired due east from a po<strong>in</strong>t on the surface of the Earth at a northern latitude with a velocity<br />

of magnitude 0 and at an <strong>in</strong>cl<strong>in</strong>ation to the horizontal of show that the lateral deflection when the projectile<br />

strikes the Earth is<br />

= 4 0<br />

3<br />

2 s<strong>in</strong> s<strong>in</strong>2 cos (12.95)<br />

where is the rotation frequency of the Earth.<br />

9. Obta<strong>in</strong> an expression for the angular deviation of a particle projected from the North Pole <strong>in</strong> a path that lies<br />

close to the surface of the earth. Is the deviation significant for a missile that makes a 4800-km flight <strong>in</strong> 10<br />

m<strong>in</strong>utes? What is the ”miss distance” if the missile is aimed directly at the target? Is the miss difference<br />

greater for a 19300-km flight at the same velocity?<br />

10. An automobile drag racer drives a car with acceleration and <strong>in</strong>stantaneous velocity . Thetiresofradius 0<br />

are not slipp<strong>in</strong>g. Derive which po<strong>in</strong>t on the tire has the greatest acceleration relative to the ground. What is<br />

this acceleration?<br />

11. Shot towers were popular <strong>in</strong> the eighteenth and n<strong>in</strong>eteenth centuries for dropp<strong>in</strong>g melted lead down tall towers<br />

to form spheres for bullets. The lead solidified while fall<strong>in</strong>g and often landed <strong>in</strong> water to cool the lead bullets.<br />

Many such shot towers were built <strong>in</strong> New York State. Assume a shot tower was constructed at latitude 42 ◦ ,<br />

and that the lead fell a distance of 27. In what direction and by how far did the lead bullets land from the<br />

direct vertical?


Chapter 13<br />

Rigid-body rotation<br />

13.1 Introduction<br />

Rigid-body rotation features prom<strong>in</strong>ently <strong>in</strong> science, eng<strong>in</strong>eer<strong>in</strong>g, and sports. Prior chapters have focussed<br />

primarily on motion of po<strong>in</strong>t particles. This chapter extends the discussion to motion of f<strong>in</strong>ite-sized rigid<br />

bodies. A rigid body is a collection of particles where the relative separations rema<strong>in</strong> rigidly fixed. In real<br />

life, there is always some motion between <strong>in</strong>dividual atoms, but usually this microscopic motion can be<br />

neglected when describ<strong>in</strong>g macroscopic properties. Note that the concept of perfect rigidity has limitations<br />

<strong>in</strong> the theory of relativity s<strong>in</strong>ce <strong>in</strong>formation cannot travel faster than the velocity of light, and thus signals<br />

cannot be transmitted <strong>in</strong>stantaneously between the ends of a rigid body which is implied if the body had<br />

perfect rigidity.<br />

The description of rigid-body rotation is most easily handled by specify<strong>in</strong>g the properties of the body<br />

<strong>in</strong> the rotat<strong>in</strong>g body-fixed coord<strong>in</strong>ate frame whereas the observables are measured <strong>in</strong> the stationary <strong>in</strong>ertial<br />

laboratory coord<strong>in</strong>ate frame. In the body-fixed coord<strong>in</strong>ate frame, the primary observable for classical<br />

mechanics is the <strong>in</strong>ertia tensor of the rigid body which is well def<strong>in</strong>ed and <strong>in</strong>dependent of the rotational<br />

motion. By contrast, <strong>in</strong> the stationary <strong>in</strong>ertial frame the observables depend sensitively on the details of the<br />

rotational motion. For example, when observed <strong>in</strong> the stationary fixed frame, rapid rotation of a long th<strong>in</strong><br />

cyl<strong>in</strong>drical pencil about the longitud<strong>in</strong>al symmetry axis gives a time-averaged shape of the pencil that looks<br />

like a th<strong>in</strong> cyl<strong>in</strong>der, whereas the time-averaged shape is a flat disk for rotation about an axis perpendicular<br />

to the symmetry axis of the pencil. In spite of this, the pencil always has the same unique <strong>in</strong>ertia tensor<br />

<strong>in</strong> the body-fixed frame. Thus the best solution for describ<strong>in</strong>g rotation of a rigid body is to use a rotation<br />

matrix that transforms from the stationary fixed frame to the <strong>in</strong>stantaneous body-fixed frame for which the<br />

moment of <strong>in</strong>ertia tensor can be evaluated. Moreover, the problem can be greatly simplified by transform<strong>in</strong>g<br />

to a body-fixed coord<strong>in</strong>ate frame that is aligned with any symmetry axes of the body s<strong>in</strong>ce then the <strong>in</strong>ertia<br />

tensor can be diagonal; this is called a pr<strong>in</strong>cipal axis system.<br />

Rigid-body rotation can be broken <strong>in</strong>to the follow<strong>in</strong>g two classifications.<br />

1) Rotation about a fixed axis:<br />

A body can be constra<strong>in</strong>ed to rotate about an axis that has a fixed location and orientation relative to<br />

the body. The h<strong>in</strong>ged door is a typical example. Rotation about a fixed axis is straightforward s<strong>in</strong>ce the<br />

axis of rotation, plus the moment of <strong>in</strong>ertia about this axis, are well def<strong>in</strong>ed and this case was discussed <strong>in</strong><br />

chapter 2127.<br />

2) Rotation about a po<strong>in</strong>t<br />

A body can be constra<strong>in</strong>ed to rotate about a fixed po<strong>in</strong>t of the body but the orientation of this rotation<br />

axis about this po<strong>in</strong>t is unconstra<strong>in</strong>ed. One example is rotation of an object fly<strong>in</strong>g freely <strong>in</strong> space which can<br />

rotate about the center of mass with any orientation. Another example is a child’s sp<strong>in</strong>n<strong>in</strong>g top which has<br />

one po<strong>in</strong>t constra<strong>in</strong>ed to touch the ground but the orientation of the rotation axis is undef<strong>in</strong>ed.<br />

The prior discussion <strong>in</strong> chapter 2127 showed that rigid-body rotation is more complicated than assumed<br />

<strong>in</strong> <strong>in</strong>troductory treatments of rigid-body rotation. It is necessary to expand the concept of moment of <strong>in</strong>ertia<br />

to the concept of the <strong>in</strong>ertia tensor, plus the fact that the angular momentum may not po<strong>in</strong>t along the<br />

rotation axis. The most general case requires consideration of rotation about a body-fixed po<strong>in</strong>t where the<br />

orientation of the axis of rotation is unconstra<strong>in</strong>ed. The concept of the <strong>in</strong>ertia tensor of a rotat<strong>in</strong>g body is<br />

309


310 CHAPTER 13. RIGID-BODY ROTATION<br />

crucial for describ<strong>in</strong>g rigid-body motion. It will be shown that work<strong>in</strong>g <strong>in</strong> the body-fixed coord<strong>in</strong>ate frame of<br />

a rotat<strong>in</strong>g body allows a description of the equations of motion <strong>in</strong> terms of the <strong>in</strong>ertia tensor for a given po<strong>in</strong>t<br />

of the body, and that it is possible to rotate the body-fixed coord<strong>in</strong>ate system <strong>in</strong>to a pr<strong>in</strong>cipal axis system<br />

where the <strong>in</strong>ertia tensor is diagonal. For any pr<strong>in</strong>cipal axis, the angular momentum is parallel to the angular<br />

velocity if it is aligned with a pr<strong>in</strong>cipal axis. The use of a pr<strong>in</strong>cipal axis system greatly simplifies treatment<br />

of rigid-body rotation and exploits the powerful and elegant matrix algebra mentioned <strong>in</strong> appendix .<br />

The follow<strong>in</strong>g discussion of rigid-body rotation is broken <strong>in</strong>to three topics, (1) the <strong>in</strong>ertia tensor of the<br />

rigid body, (2) the transformation between the rotat<strong>in</strong>g body-fixed coord<strong>in</strong>ate system and the laboratory<br />

frame, i.e., the Euler angles specify<strong>in</strong>g the orientation of the body-fixed coord<strong>in</strong>ate frame with respect to the<br />

laboratory frame, and (3) Lagrange and Euler’s equations of motion for rigid-bodies. This is followed by a<br />

discussion of practical applications.<br />

13.2 Rigid-body coord<strong>in</strong>ates<br />

Motion of a rigid body is a special case for motion of the -body system when the relative positions of<br />

the bodies are related. It was shown <strong>in</strong> chapter 2 that the motion of a rigid body can be broken <strong>in</strong>to<br />

a comb<strong>in</strong>ation of a l<strong>in</strong>ear translation of some po<strong>in</strong>t <strong>in</strong> the body, plus rotation of the body about an axis<br />

through that po<strong>in</strong>t. This is called Chasles’ Theorem. Thus the position of every particle <strong>in</strong> the rigid body<br />

is fixed with respect to one po<strong>in</strong>t <strong>in</strong> the body. If the fixed po<strong>in</strong>t of the body is chosen to be the center of<br />

mass, then, as discussed <strong>in</strong> chapter 2, it is possible to separate the k<strong>in</strong>etic energy, l<strong>in</strong>ear momentum, and<br />

angular momentum <strong>in</strong>to the center-of-mass motion, plus the motion about the center of mass. Thus the<br />

behavior of the body can be described completely us<strong>in</strong>g only six <strong>in</strong>dependent coord<strong>in</strong>ates governed by six<br />

equations of motion, three for translation and three for rotation.<br />

Referred to an <strong>in</strong>ertial frame, the translational motion of the center of mass is governed by<br />

F = P<br />

(13.1)<br />

<br />

while the rotational motion about the center of mass is determ<strong>in</strong>ed by<br />

N = L<br />

(13.2)<br />

<br />

where the external force F and external torque N are identified separately from the <strong>in</strong>ternal forces act<strong>in</strong>g<br />

between the particles <strong>in</strong> the rigid body. It will be assumed that the <strong>in</strong>ternal forces are central and thus do<br />

not contribute to the angular momentum.<br />

The location of any fixed po<strong>in</strong>t <strong>in</strong> the body, such as the center of mass, can be specified by three generalized<br />

cartesian coord<strong>in</strong>ates with respect to a fixed frame. The rotation of the body-fixed axis system about this<br />

fixed po<strong>in</strong>t <strong>in</strong> the body can be described <strong>in</strong> terms of three <strong>in</strong>dependent angles with respect to the fixed frame.<br />

There are several possible sets of orthogonal angles that can be used to describe the rotation. This book<br />

uses the Euler angles which correspond first to a rotation about the -axis, then a rotation about<br />

the axis subsequent to the first rotation, and f<strong>in</strong>ally a rotation about the new axis follow<strong>in</strong>g the first<br />

two rotations. The Euler angles will be discussed <strong>in</strong> detail follow<strong>in</strong>g <strong>in</strong>troduction of the <strong>in</strong>ertia tensor and<br />

angular momentum.<br />

13.3 Rigid-body rotation about a body-fixed po<strong>in</strong>t<br />

With respect to some po<strong>in</strong>t fixed <strong>in</strong> the body coord<strong>in</strong>ate system, the angular momentum of the body is<br />

given by<br />

X X<br />

L = L = r × p (13.3)<br />

<br />

There are two especially convenient choices for the fixed po<strong>in</strong>t . If no po<strong>in</strong>t <strong>in</strong> the body is fixed with<br />

respect to an <strong>in</strong>ertial coord<strong>in</strong>ate system, then it is best to choose as the center of mass. If one po<strong>in</strong>t of<br />

the body is fixed with respect to a fixed <strong>in</strong>ertial coord<strong>in</strong>ate system, such as a po<strong>in</strong>t on the ground where a<br />

child’s sp<strong>in</strong>n<strong>in</strong>g top touches, then it is best to choose this stationary po<strong>in</strong>t as the body-fixed po<strong>in</strong>t


13.3. RIGID-BODYROTATIONABOUTABODY-FIXEDPOINT 311<br />

Consider a rigid body composed of particles of mass<br />

where =1 2 3 As discussed <strong>in</strong> chapter 124, ifthe<br />

body rotates with an <strong>in</strong>stantaneous angular velocity ω about<br />

some fixed po<strong>in</strong>t, with respect to the body-fixed coord<strong>in</strong>ate<br />

system, and this po<strong>in</strong>t has an <strong>in</strong>stantaneous translational velocity<br />

V with respect to the fixed (<strong>in</strong>ertial) coord<strong>in</strong>ate system,<br />

see figure 131, then the <strong>in</strong>stantaneous velocity v of the <br />

particle <strong>in</strong> the fixed frame of reference is given by<br />

v = V + v 00 + ω × r 0 (13.4)<br />

However, for a rigid body, the velocity of a body-fixed po<strong>in</strong>t<br />

with respect to the body is zero, that is v 00 =0 thus<br />

v = V + ω × r 0 (13.5)<br />

Consider the translational velocity of the body-fixed po<strong>in</strong>t<br />

to be zero, i.e. V =0 and let R =0 then r = r 0 .These<br />

assumptions allow the l<strong>in</strong>ear momentum of the particle to<br />

be written as<br />

<br />

p = v = ω × r (13.6)<br />

Therefore<br />

X<br />

X<br />

L = r × p = r × (ω × r ) (13.7)<br />

Us<strong>in</strong>g the vector identity<br />

leads to<br />

<br />

A × (B × A) = 2 B − A (A · B)<br />

L =<br />

Figure 13.1: Inf<strong>in</strong>itessimal displacement 0 <br />

<strong>in</strong> the primed frame, broken <strong>in</strong>to a part <br />

due to rotation of the primed frame plus a<br />

part 00 due to displacement with respect to<br />

this rotat<strong>in</strong>g frame.<br />

X £<br />

<br />

2<br />

ω − r (r · ω) ¤ (13.8)<br />

<br />

The angular momentum can be expressed <strong>in</strong> terms of components of ω and r 0 relative to the body-fixed<br />

frame. The follow<strong>in</strong>g formulae can be written more compactly if r =( ) <strong>in</strong> the rotat<strong>in</strong>g body-fixed<br />

frame,iswritten<strong>in</strong>theformr =( 1 2 3 ) where the axes are def<strong>in</strong>ed by the numbers 1 2 3 rather<br />

than . In this notation, the angular momentum is written <strong>in</strong> component form as<br />

⎡<br />

⎛ ⎞⎤<br />

X X<br />

= <br />

⎣ 2 − <br />

⎝ X <br />

⎠⎦ (13.9)<br />

<br />

<br />

<br />

Assume the Kronecker delta relation<br />

3X<br />

= (13.10)<br />

where<br />

<br />

= 1 = <br />

= 0 6= <br />

Substitute (1310) <strong>in</strong> (139) gives<br />

=<br />

=<br />

"<br />

#<br />

X X X<br />

2 − <br />

<br />

<br />

"<br />

3X <br />

Ã<br />

!#<br />

X X<br />

2 − <br />

<br />

<br />

(13.11)


312 CHAPTER 13. RIGID-BODY ROTATION<br />

13.4 Inertia tensor<br />

Thesquarebracketterm<strong>in</strong>(1311) is called the moment of <strong>in</strong>ertia tensor I which is usually referred to<br />

as the <strong>in</strong>ertia tensor<br />

" Ã<br />

X<br />

3X<br />

! #<br />

≡ 2 − (13.12)<br />

<br />

<br />

In most cases it is more useful to express the components of the <strong>in</strong>ertia tensor <strong>in</strong> an <strong>in</strong>tegral form over<br />

the mass distribution rather than a summation for discrete bodies. That is,<br />

Z Ã Ã 3X<br />

! !<br />

= (r 0 ) 2 − (13.13)<br />

<br />

The <strong>in</strong>ertia tensor is easier to understand when written <strong>in</strong> cartesian coord<strong>in</strong>ates r 0 =( ) rather<br />

than <strong>in</strong> the form r 0 =( 1 2 3 ) Then, the diagonal moments of <strong>in</strong>ertia of the <strong>in</strong>ertia tensor are<br />

<br />

<br />

<br />

≡<br />

≡<br />

≡<br />

X £<br />

<br />

2<br />

+ 2 + 2 − 2 ¤ X<br />

£<br />

= <br />

2<br />

+ <br />

2 ¤<br />

<br />

X £<br />

<br />

2<br />

+ 2 + 2 − <br />

2 ¤ X<br />

£<br />

= <br />

2<br />

+ <br />

2 ¤<br />

<br />

X £<br />

<br />

2<br />

+ 2 + 2 − <br />

2 ¤ X <br />

£<br />

= <br />

2<br />

+ <br />

2 ¤<br />

<br />

<br />

<br />

<br />

(13.14)<br />

while the off-diagonal products of <strong>in</strong>ertia are<br />

= ≡−<br />

= ≡−<br />

= ≡−<br />

X<br />

[ ] (13.15)<br />

<br />

X<br />

[ ]<br />

<br />

X<br />

[ ]<br />

<br />

Note that the products of <strong>in</strong>ertia are symmetric <strong>in</strong> that<br />

= (13.16)<br />

The above notation for the <strong>in</strong>ertia tensor allows the angular momentum (1312) to be written as<br />

=<br />

3X<br />

(13.17)<br />

<br />

Expanded <strong>in</strong> cartesian coord<strong>in</strong>ates<br />

= + + (13.18)<br />

= + + <br />

= + + <br />

Note that every fixed po<strong>in</strong>t <strong>in</strong> a body has a specific <strong>in</strong>ertia tensor. The components of the <strong>in</strong>ertia tensor<br />

at a specified po<strong>in</strong>t depend on the orientation of the coord<strong>in</strong>ate frame whose orig<strong>in</strong> is located at the specified<br />

fixed po<strong>in</strong>t. For example, the <strong>in</strong>ertia tensor for a cube is very different when the fixed po<strong>in</strong>t is at the center<br />

of mass compared with when the fixed po<strong>in</strong>t is at a corner of the cube.


13.5. MATRIX AND TENSOR FORMULATIONS OF RIGID-BODY ROTATION 313<br />

13.5 Matrix and tensor formulations of rigid-body rotation<br />

The prior notation is clumsy and can be streaml<strong>in</strong>ed by use of matrix methods. Write the <strong>in</strong>ertia tensor <strong>in</strong><br />

a matrix form as<br />

⎛<br />

{I} = ⎝ ⎞<br />

11 12 13<br />

21 22 23<br />

⎠ (13.19)<br />

31 32 33<br />

The angular velocity and angular momentum both can be written as a column vectors, that is<br />

⎛<br />

ω = ⎝ ⎞<br />

⎛<br />

1<br />

2<br />

⎠ L = ⎝ ⎞<br />

1<br />

2<br />

⎠ (13.20)<br />

3<br />

3<br />

As discussed <strong>in</strong> appendix 2, equation(1318) nowcanbewritten<strong>in</strong>tensornotationasan<strong>in</strong>nerproduct<br />

of the form<br />

L = {I}·ω (13.21)<br />

Note that the above notation uses boldface for the <strong>in</strong>ertia tensor I, imply<strong>in</strong>g a rank-2 tensor representation,<br />

while the angular velocity ω and the angular momentum L are written as column vectors. The <strong>in</strong>ertia tensor<br />

is a 9-component rank-2 tensor def<strong>in</strong>ed as the ratio of the angular momentum vector L and the angular<br />

velocity ω.<br />

{I} = L ω<br />

(13.22)<br />

Note that, as described <strong>in</strong> appendix , the <strong>in</strong>ner product of a vector ω, whichistherank1 tensor, and a<br />

rank 2 tensor {I} leads to the vector L. This compact notation exploits the fact that the matrix and tensor<br />

representation are completely equivalent, and are ideally suited to the description of rigid-body rotation.<br />

13.6 Pr<strong>in</strong>cipal axis system<br />

The <strong>in</strong>ertia tensor is a real symmetric matrix because of the symmetry given by equation (1316) A property<br />

of real symmetric matrices is that there exists an orientation of the coord<strong>in</strong>ate frame, with its orig<strong>in</strong> at the<br />

chosen body-fixed po<strong>in</strong>t such that the <strong>in</strong>ertia tensor is diagonal. The coord<strong>in</strong>ate system for which the<br />

<strong>in</strong>ertia tensor is diagonal is called the Pr<strong>in</strong>cipal axis system which has three perpendicular pr<strong>in</strong>cipal<br />

axes. Thus, <strong>in</strong> the pr<strong>in</strong>cipal axis system, the <strong>in</strong>ertia tensor has the form<br />

⎛<br />

{I} = ⎝ ⎞<br />

11 0 0<br />

0 22 0 ⎠ (13.23)<br />

0 0 33<br />

where are real numbers, which are called the pr<strong>in</strong>cipal moments of <strong>in</strong>ertia of the body, and are<br />

usually written as . When the angular velocity vector ω po<strong>in</strong>ts along any pr<strong>in</strong>cipal axis unit vector ˆ, then<br />

the angular momentum L is parallel to ω and the magnitude of the pr<strong>in</strong>cipal moment of <strong>in</strong>ertia about this<br />

pr<strong>in</strong>cipal axis is given by the relation<br />

ˆ = ˆ (13.24)<br />

The pr<strong>in</strong>cipal axes are fixed relative to the shape of the rigid body and they are <strong>in</strong>variant to the orientation<br />

of the body-fixed coord<strong>in</strong>ate system used to evaluate the <strong>in</strong>ertia tensor. The advantage of hav<strong>in</strong>g the bodyfixed<br />

coord<strong>in</strong>ate frame aligned with the pr<strong>in</strong>cipal axis coord<strong>in</strong>ate frame is that then the <strong>in</strong>ertia tensor is<br />

diagonal, which greatly simplifies the matrix algebra. Even when the body-fixed coord<strong>in</strong>ate system is not<br />

aligned with the pr<strong>in</strong>cipal axis frame, if the angular velocity is specified to po<strong>in</strong>t along a pr<strong>in</strong>cipal axis then<br />

the correspond<strong>in</strong>g moment of <strong>in</strong>ertia will be given by (1324) <br />

In pr<strong>in</strong>ciple it is possible to locate the pr<strong>in</strong>cipal axes by vary<strong>in</strong>g the orientation of the angular velocity<br />

vector ω to f<strong>in</strong>d those orientations for which the angular momentum L and angular velocity ω are parallel<br />

which characterizes the pr<strong>in</strong>cipal axes. However, the best approach is to diagonalize the <strong>in</strong>ertia tensor.


314 CHAPTER 13. RIGID-BODY ROTATION<br />

13.7 Diagonalize the <strong>in</strong>ertia tensor<br />

F<strong>in</strong>d<strong>in</strong>g the three pr<strong>in</strong>cipal axes <strong>in</strong>volves diagonaliz<strong>in</strong>g the <strong>in</strong>ertia tensor, which is the classic eigenvalue<br />

problem discussed <strong>in</strong> appendix . Solution of the eigenvalue problem for rigid-body motion corresponds to<br />

a rotation of the coord<strong>in</strong>ate frame to the pr<strong>in</strong>cipal axes result<strong>in</strong>g <strong>in</strong> the matrix<br />

{I}·ω = ω (13.25)<br />

where comprises the three-valued eigenvalues, while the correspond<strong>in</strong>g vector ω is the eigenvector. Appendix<br />

4 gives the solution of the matrix relation<br />

{I}·ω = {I} ω (13.26)<br />

where are three-valued eigen values for the pr<strong>in</strong>cipal axis moments of <strong>in</strong>ertia, and {I} is the unity tensor,<br />

equation 24.<br />

⎧ ⎫<br />

⎨ 1 0 0 ⎬<br />

{I} ≡ 0 1 0<br />

(13.27)<br />

⎩ ⎭<br />

0 0 1<br />

Rewrit<strong>in</strong>g (1326) gives<br />

({I} − {I}) · ω =0 (13.28)<br />

This is a matrix equation of the form A · ω =0 where A is a 3 × 3 matrix and ω is a vector with values<br />

The matrix equation A · ω =0 really corresponds to three simultaneous equations for the three<br />

numbers . It is a well-known property of equations like (1328) that they have a non-zero solution<br />

if, and only if, the determ<strong>in</strong>ant det(A) is zero, that is<br />

det(I−I)=0 (13.29)<br />

This is called the characteristic equation, or secular equation for the matrix I. The determ<strong>in</strong>ant<br />

<strong>in</strong>volved is a cubic equation <strong>in</strong> the value of that gives the three pr<strong>in</strong>cipal moments of <strong>in</strong>ertia. Insert<strong>in</strong>g<br />

one of the three values of <strong>in</strong>to equation (1317) gives the correspond<strong>in</strong>g eigenvector . Apply<strong>in</strong>g the above<br />

eigenvalue problem to rigid-body rotation corresponds to requir<strong>in</strong>g that some arbitrary set of body-fixed<br />

axes be the pr<strong>in</strong>cipal axes of <strong>in</strong>ertia. This is obta<strong>in</strong>ed by rotat<strong>in</strong>g the body-fixed axis system such that<br />

or<br />

1 = 11 1 + 12 2 + 13 3 = 1 (13.30)<br />

2 = 21 1 + 22 2 + 23 3 = 2<br />

3 = 31 1 + 32 2 + 33 3 = 3<br />

( 11 − ) 1 + 12 2 + 13 3 = 0 (13.31)<br />

21 1 +( 22 − ) 2 + 23 3 = 0<br />

31 1 + 32 2 +( 33 − ) 3 = 0<br />

These equations have a non-trivial solution for the ratios 1 : 2 : 3 s<strong>in</strong>ce the determ<strong>in</strong>ant vanishes, that is<br />

( 11 − ) 12 13<br />

21 ( 22 − ) 23<br />

=0 (13.32)<br />

¯ 31 32 ( 33 − ) ¯<br />

The expansion of this determ<strong>in</strong>ant leads to a cubic equation with three roots for This is the secular<br />

equation for whose eigenvalues are the pr<strong>in</strong>cipal moments of <strong>in</strong>ertia.<br />

The directions of the pr<strong>in</strong>cipal axes, that is the eigenvectors, can be found by substitut<strong>in</strong>g the correspond<strong>in</strong>g<br />

solution for <strong>in</strong>to the prior equation. Thus for eigensolution 1 the eigenvector is given by<br />

solv<strong>in</strong>g<br />

( 11 − 1 ) 11 + 12 21 + 13 31 = 0 (13.33)<br />

21 11 +( 22 − 1 ) 21 + 23 31 = 0<br />

31 11 + 32 21 +( 33 − 1 ) 31 = 0


13.8. PARALLEL-AXIS THEOREM 315<br />

These equations are solved for the ratios 11 : 21 : 31 which are the direction numbers of the pr<strong>in</strong>ciple axis<br />

system correspond<strong>in</strong>g to solution 1 This pr<strong>in</strong>cipal axis system is def<strong>in</strong>ed relative to the orig<strong>in</strong>al coord<strong>in</strong>ate<br />

system. This procedure is repeated to f<strong>in</strong>d the orientation of the other two mutually perpendicular pr<strong>in</strong>cipal<br />

axes.<br />

13.8 Parallel-axis theorem<br />

The values of the components of the <strong>in</strong>ertia tensor depend on both the<br />

location and the orientation about which the body rotates relative to<br />

the body-fixed coord<strong>in</strong>ate system. The parallel-axis theorem is valuable<br />

for relat<strong>in</strong>g the <strong>in</strong>ertia tensor for rotation about parallel axes pass<strong>in</strong>g<br />

through different po<strong>in</strong>ts fixed with respect to the rigid body. For example,<br />

one may wish to relate the <strong>in</strong>ertia tensor through the center of<br />

mass to another location that is constra<strong>in</strong>ed to rema<strong>in</strong> stationary, like<br />

the tip of the sp<strong>in</strong>n<strong>in</strong>g top.<br />

Consider the mass at the location r =( 1 2 3 ) with respect<br />

to the orig<strong>in</strong> of the center of mass body-fixed coord<strong>in</strong>ate system .<br />

Transform to an arbitrary but parallel body-fixed coord<strong>in</strong>ate system<br />

, that is, the coord<strong>in</strong>ate axes have the same orientation as the center<br />

of mass coord<strong>in</strong>ate system. The location of the mass with respect<br />

to this arbitrary coord<strong>in</strong>ate system is R =( 1 2 3 ) That is, the<br />

general vectors for the two coord<strong>in</strong>ates systems are related by<br />

X1<br />

Q<br />

X<br />

3<br />

a<br />

x3<br />

O<br />

x<br />

1<br />

R<br />

r<br />

x2<br />

X<br />

2<br />

R = a + r (13.34)<br />

where a is the vector connect<strong>in</strong>g the orig<strong>in</strong>s of the coord<strong>in</strong>ate systems<br />

and illustrated <strong>in</strong> figure 132. The elements of the <strong>in</strong>ertia tensor<br />

with respect to axis system are given by equation 1312 to be<br />

" Ã<br />

X<br />

3X<br />

≡ <br />

2<br />

<br />

<br />

!<br />

− <br />

#<br />

Figure 13.2: Transformation between<br />

two parallel body-coord<strong>in</strong>ate<br />

systems, O and Q.<br />

The components along the three axes for each of the two coord<strong>in</strong>ate systems are related by<br />

(13.35)<br />

= + (13.36)<br />

Substitut<strong>in</strong>g these <strong>in</strong>to the above <strong>in</strong>ertia tensor relation gives<br />

" Ã<br />

X<br />

3X<br />

! #<br />

= ( + ) 2 − ( + )( + )<br />

(13.37)<br />

<br />

<br />

" Ã<br />

X<br />

3X<br />

! # " Ã<br />

X<br />

3X<br />

= 2 ¡<br />

− + 2 + 2 ¢ ! #<br />

− ( + + )<br />

<br />

<br />

<br />

<br />

The first summation on the right-hand side corresponds to the elements of the <strong>in</strong>ertia tensor <strong>in</strong> the<br />

center-of-mass frame. Thus the terms can be regrouped to give<br />

Ã<br />

! "<br />

#<br />

X<br />

3X<br />

X<br />

3X<br />

≡ + 2 − + − − (13.38)<br />

<br />

<br />

<br />

<br />

However, each term <strong>in</strong> the last bracket <strong>in</strong>volves a sum of the form P <br />

Takethecoord<strong>in</strong>atesystem<br />

to be with respect to the center of mass for which<br />

2 <br />

<br />

X<br />

r 0 =0 (13.39)


316 CHAPTER 13. RIGID-BODY ROTATION<br />

This also applies to each component , thatis<br />

X<br />

=0 (13.40)<br />

Therefore all of the terms <strong>in</strong> the last bracket cancel leav<strong>in</strong>g<br />

Ã<br />

!<br />

X<br />

3X<br />

≡ + 2 − <br />

<br />

<br />

<br />

<br />

(13.41)<br />

But P <br />

= and P 3<br />

2 = 2 thus<br />

≡ + ¡ 2 − <br />

¢<br />

(13.42)<br />

where is the center-of-mass <strong>in</strong>ertia tensor. This is the general form of Ste<strong>in</strong>er’s parallel-axis theorem.<br />

As an example, the moment of <strong>in</strong>ertia around the 1 axis is given by<br />

11 ≡ 11 + ¡¡ 2 1 + 2 2 + 2 3¢<br />

11 − 2 1¢<br />

= 11 + ¡ 2 2 + 2 ¢<br />

3<br />

(13.43)<br />

which corresponds to the elementary statement that the difference <strong>in</strong> the moments of <strong>in</strong>ertia equals the<br />

massofthebodymultipliedbythesquareofthedistancebetweentheparallelaxes, 1 1 Note that the<br />

m<strong>in</strong>imum moment of <strong>in</strong>ertia of a body is which is about the center of mass.<br />

13.1 Example: Inertia tensor of a solid cube rotat<strong>in</strong>g about the center of mass.<br />

The complicated expressions for the <strong>in</strong>ertia tensor can be understood<br />

us<strong>in</strong>g the example of a uniform solid cube with side ,<br />

density and mass = 3 rotat<strong>in</strong>g about different axes. Assume<br />

that the orig<strong>in</strong> of the coord<strong>in</strong>ate system is at the center<br />

of mass with the axes perpendicular to the centers of the faces of<br />

the cube.<br />

The components of the <strong>in</strong>ertia tensor can be calculated us<strong>in</strong>g<br />

(1313) written as an <strong>in</strong>tegral over the mass distribution rather<br />

than a summation.<br />

Z Ã Ã 3X<br />

! !<br />

= (r 0 ) 2 − <br />

<br />

O<br />

Thus<br />

Z 2 Z 2 Z 2<br />

¡<br />

11 = <br />

<br />

2<br />

2 + 2 ¢<br />

3 3 2 1<br />

−2 −2 −2<br />

= 1 6 5 = 1 6 2 = 22 = 33<br />

Inertia tensor of a uniform solid cube of<br />

side about the center of mass and a<br />

corner of the cube . The vector is the<br />

vector distance between and .<br />

By symmetry the diagonal moments of <strong>in</strong>ertia about each face<br />

are identical. Similarly the products of <strong>in</strong>ertia are given by<br />

12 = −<br />

Z 2 Z 2 Z 2<br />

−2<br />

−2<br />

−2<br />

( 1 2 ) 3 2 1 =0<br />

Thus the <strong>in</strong>ertia tensor is given by<br />

⎛<br />

I = 1 6 2 ⎝ 1 0 0 ⎞<br />

0 1 0 ⎠<br />

0 0 1<br />

Note that this <strong>in</strong>ertia tensor is diagonal imply<strong>in</strong>g that this is the pr<strong>in</strong>cipal axis system. In this case all three<br />

pr<strong>in</strong>cipal moments of <strong>in</strong>ertia are identical and perpendicular to the centers of the faces of the cube. This is<br />

as expected from the symmetry of the cubic geometry.


13.8. PARALLEL-AXIS THEOREM 317<br />

13.2 Example: Inertia tensor of about a corner of a solid cube.<br />

a) Direct calculation Let one corner of the cube be the orig<strong>in</strong> of the coord<strong>in</strong>ate system and assume<br />

that the three adjacent sides of the cube lie along the coord<strong>in</strong>ate axes. The components of the <strong>in</strong>ertia tensor<br />

can be calculated us<strong>in</strong>g (1313) Thus<br />

11 = <br />

12 = −<br />

Z Z Z <br />

0 0 0<br />

Z Z Z <br />

Thus, evaluat<strong>in</strong>g all the n<strong>in</strong>e components gives<br />

0<br />

0<br />

0<br />

¡<br />

<br />

2<br />

2 + 2 3¢<br />

3 2 1 = 2 3 5 = 2 3 2<br />

⎛<br />

I = 1<br />

12 2 ⎝<br />

( 1 2 ) 3 2 1 = − 1 4 5 = − 1 4 2<br />

8 −3 −3<br />

−3 8 −3<br />

−3 −3 8<br />

b) Parallel-axis theorem This <strong>in</strong>ertia tensor also can be calculated us<strong>in</strong>g the parallel-axis theorem to<br />

relate the moment of <strong>in</strong>ertia about the corner, to that at the center of mass. As shown <strong>in</strong> the figure, the<br />

vector has components<br />

1 = 2 = 3 = 2<br />

Apply<strong>in</strong>g the parallel-axis theorem gives<br />

11 = 11 + ¡ 2 − 2 1¢<br />

= 11 + ¡ 2 2 + 2 ¢ 1<br />

3 =<br />

6 2 + 1 2 2 = 2 3 2<br />

and similarly for 22 and 33 . The off-diagonal terms are given by<br />

12 = 12 + (− 1 2 )=− 1 4 2<br />

Thus the <strong>in</strong>ertia tensor, transposed from the center of mass, to the corner of the cube is<br />

⎛<br />

2<br />

I 3<br />

= ⎝<br />

2 − 1 4 2 − 1 ⎞ ⎛<br />

⎞<br />

4 2<br />

− 1 4 2 2 3 2 − 1 4 2 ⎠ = 1 8 −3 −3<br />

− 1 4 2 − 1 4 2 2 12 2 ⎝ −3 8 −3 ⎠<br />

3 2 −3 −3 8<br />

This <strong>in</strong>ertia tensor about the corner of the cube, is the same as that obta<strong>in</strong>ed by direct <strong>in</strong>tegration.<br />

c) Pr<strong>in</strong>cipal moments of <strong>in</strong>ertia The coord<strong>in</strong>ate axis frame used for rotation about the corner of the<br />

cube is not a pr<strong>in</strong>cipal axis frame. Therefore let us diagonalize the <strong>in</strong>ertia tensor to f<strong>in</strong>d the pr<strong>in</strong>cipal<br />

axis frame the pr<strong>in</strong>cipal moments of <strong>in</strong>ertia about a corner. To achieve this requires solv<strong>in</strong>g the secular<br />

determ<strong>in</strong>ant<br />

¡ 2<br />

3 2 − ¢ − 1 4 2 − 1 4 2<br />

⎞<br />

⎠<br />

¡<br />

−<br />

¯¯¯¯¯¯<br />

1 2<br />

4 2 3 2 − ¢ ¡<br />

− 1 4 2<br />

− 1 4 2 − 1 2<br />

4 2 3 2 − <br />

¯¯¯¯¯¯ ¢ =0<br />

The value of a determ<strong>in</strong>ant is not affected by add<strong>in</strong>g or subtract<strong>in</strong>g any row or column from any other<br />

row or column. Subtract row 1 from row 2 gives<br />

¡ 2 3 2 − ¢ ¡ − 1 4 2 − 1 4 2<br />

− 11<br />

12<br />

¯<br />

2 + 11<br />

12 2 − ¢ ¡ 0<br />

− 1 4 2 − 1 2<br />

4 2 3 2 − <br />

¯¯¯¯¯¯ ¢ =0<br />

The determ<strong>in</strong>ant of this matrix is straightforward to evaluate and equals<br />

µ µ µ <br />

1 11<br />

11<br />

6 2 − <br />

12 2 − <br />

12 2 − =0


318 CHAPTER 13. RIGID-BODY ROTATION<br />

Thus the roots are<br />

⎛<br />

⎞<br />

1<br />

I 6<br />

= ⎝<br />

2 0 0<br />

11<br />

0<br />

12 2 0 ⎠<br />

11<br />

0 0<br />

12 2<br />

The identical roots 22 = 33 = 11<br />

12 2 imply that the pr<strong>in</strong>cipal axis associated with 11 must be a symmetry<br />

axis. The orientation can be found by substitut<strong>in</strong>g 11 <strong>in</strong>to the above equation<br />

⎛<br />

({I} − {I}) · ω = 1<br />

12 2 ⎝<br />

6 −3 −3<br />

−3 6 −3<br />

−3 −3 6<br />

⎞ ⎛<br />

⎠ ⎝ ⎞<br />

11<br />

21<br />

⎠ =0<br />

31<br />

where the second subscript 1 attached to signifies that this solution corresponds to 11 This gives<br />

2 11 − 21 − 31 = 0<br />

− 11 +2 21 − 31 = 0<br />

− 11 − 21 +2 31 = 0<br />

Solv<strong>in</strong>g<br />

⎛<br />

these three equations gives the unit vector for the first pr<strong>in</strong>cipal axis for which 11 = 1 6 2 to be<br />

ê 1 = √ 1 ⎝ 1 ⎞<br />

3<br />

1 ⎠. This can be repeated to f<strong>in</strong>d the other two pr<strong>in</strong>cipal axes by substitut<strong>in</strong>g 22 = 11<br />

12 2 This<br />

1<br />

gives for the second pr<strong>in</strong>cipal moment 22<br />

⎛<br />

⎞ ⎛<br />

({I} − {I}) · ω = 1 −3 −3 −3<br />

12 2 ⎝ −3 −3 −3 ⎠ ⎝ ⎞<br />

12<br />

22<br />

⎠ =0<br />

−3 −3 −3 32<br />

This results <strong>in</strong> three identical equations for the components of butallthreeequationsarethesame,namely<br />

12 + 22 + 32 =0<br />

This does not uniquely determ<strong>in</strong>e the direction of However, it does imply that 2 correspond<strong>in</strong>g to the<br />

second pr<strong>in</strong>cipal axis has the property that<br />

ˆω · ê 1 =0<br />

that is, any direction of ˆ 2 that is perpendicular to ˆ 1 is acceptable. In other words; any two orthogonal unit<br />

vectors ˆ 2 and ˆ 3 that are perpendicular to ˆ 1 are acceptable. This ambiguity exists whenever two eigenvalues<br />

are equal; the three pr<strong>in</strong>cipal axes are only uniquely def<strong>in</strong>ed if all three eigenvalues are different. The same<br />

ambiguity exist when all three eigenvalues are identical as occurs for the pr<strong>in</strong>cipal moments of <strong>in</strong>ertia about<br />

the center-of-mass of a uniform solid cube. This expla<strong>in</strong>s why the pr<strong>in</strong>cipal moment of <strong>in</strong>ertia for the diagonal<br />

of the cube, that passes through the center of mass, has the same moment as when the pr<strong>in</strong>cipal axes pass<br />

through the center of the faces of the cube.<br />

13.9 Perpendicular-axis theorem for plane lam<strong>in</strong>ae<br />

Rigid-body rotation of th<strong>in</strong> plane lam<strong>in</strong>ae objects is encountered frequently. Examples of such lam<strong>in</strong>ae<br />

bodies are a plane sheet of metal, a th<strong>in</strong> door, a bicycle wheel, a th<strong>in</strong> envelope or book. Deriv<strong>in</strong>g the <strong>in</strong>ertia<br />

tensor for a plane lam<strong>in</strong>a is relatively simple because there are limits on the possible relative magnitude<br />

of the pr<strong>in</strong>cipal moments of <strong>in</strong>ertia. Consider that the pr<strong>in</strong>cipal axis are along the coord<strong>in</strong>ate axes.<br />

Then the sum of two pr<strong>in</strong>cipal moments of <strong>in</strong>ertia about the center of mass are<br />

Z<br />

Z<br />

+ = ( 2 + 2 ) + ( 2 + 2 )<br />

Z<br />

Z Z<br />

= ( 2 + 2 ) +2 2 ≥ ( 2 + 2 ) = (13.44)


13.10. GENERAL PROPERTIES OF THE INERTIA TENSOR 319<br />

Note that for any body the three pr<strong>in</strong>cipal moments of <strong>in</strong>ertia must satisfy the triangle rule that the sum of<br />

any pair must exceed or equal the third. Moreover, if the body is a th<strong>in</strong> lam<strong>in</strong>a with thickness =0 that<br />

is, a th<strong>in</strong> plate <strong>in</strong> the − plane, then<br />

+ = (13.45)<br />

This perpendicular-axis theorem can be very useful for solv<strong>in</strong>g problems <strong>in</strong>volv<strong>in</strong>g rotation of plane lam<strong>in</strong>ae.<br />

The opposite of a plane lam<strong>in</strong>ae is a long th<strong>in</strong> cyl<strong>in</strong>drical needle of mass , length, and radius <br />

Along the symmetry axis the pr<strong>in</strong>cipal moments are = 1 2 2 → 0 as → 0 while perpendicular to the<br />

symmetry axis = = 1 12 2 . These satisfy the triangle rule.<br />

13.3 Example: Inertia tensor of a hula hoop<br />

The hula hoop is a th<strong>in</strong> plane circular r<strong>in</strong>g or radius and mass . Assume that the symmetry axis of<br />

the circular r<strong>in</strong>g is the 3 axis.<br />

a) The pr<strong>in</strong>cipal moments of <strong>in</strong>ertia about the center of mass: The pr<strong>in</strong>cipal moment of <strong>in</strong>ertia along the<br />

3 axis is 33 = 2 . Then equation 1345 plus symmetry tells us that the two pr<strong>in</strong>cipal moments of <strong>in</strong>ertia<br />

<strong>in</strong> the plane of the hula hoop must be 11 = 22 = 1 2 2 .<br />

b) The pr<strong>in</strong>cipal moments of <strong>in</strong>ertia about the periphery of the r<strong>in</strong>g: Us<strong>in</strong>g the Parallel-axis theorem<br />

tells us that the moment perpendicular to the plane of the hula hoop 33 =2 2 . In the plane of the hoop<br />

the moment tangential to the hoop is 11 = 3 2 2 and the moment radial to the hoop 22 = 1 2 2 . The<br />

hula dancer often sw<strong>in</strong>gs the hoop about the periphery and perpendicular to the plane by sw<strong>in</strong>g<strong>in</strong>g their hips.<br />

Another movement is jump<strong>in</strong>g through the hoop by rotat<strong>in</strong>g the hoop tangential to the periphery. Calculation<br />

of such maneuvers requires knowledge of these pr<strong>in</strong>cipal moments of <strong>in</strong>ertia.<br />

13.4 Example: Inertia tensor of a th<strong>in</strong> book<br />

Consider a th<strong>in</strong> rectangular book of mass width and length with thickness and .<br />

About the center of mass the <strong>in</strong>ertia tensor perpendicular to the plane of the book is 33 = 12 (2 + 2 ). The<br />

other two moments are 11 = 12 2 and 22 = 12 2 which satisfy equation 1345.<br />

13.10 General properties of the <strong>in</strong>ertia tensor<br />

13.10.1 Inertial equivalence<br />

The elements of the <strong>in</strong>ertia tensor, the values of the pr<strong>in</strong>cipal moments of <strong>in</strong>ertia, and the orientation of the<br />

pr<strong>in</strong>cipal axes for a rigid body, all depend on the choice of orig<strong>in</strong> for the system. Recall that for the k<strong>in</strong>etic<br />

energy to be separable <strong>in</strong>to translational and rotational portions, the orig<strong>in</strong> of the body coord<strong>in</strong>ate system<br />

must co<strong>in</strong>cide with the center of mass of the body. However, for any choice of the orig<strong>in</strong> of any body, there<br />

always exists an orientation of the axes that diagonalizes the <strong>in</strong>ertia tensor.<br />

The <strong>in</strong>ertial properties of a body for rotation about a specific body-fixed location is def<strong>in</strong>ed completely<br />

by only three pr<strong>in</strong>cipal moments of <strong>in</strong>ertia irrespective of the detailed shape of the body. As a result, the<br />

<strong>in</strong>ertial properties of any body about a body-fixed po<strong>in</strong>t are equivalent to that of an ellipsoid that has the<br />

same three pr<strong>in</strong>cipal moments of <strong>in</strong>ertia. The symmetry properties of this equivalent ellipsoidal body def<strong>in</strong>e<br />

the symmetry of the <strong>in</strong>ertial properties of the body. If a body has some simple symmetry then usually it is<br />

obvious as to what will be the pr<strong>in</strong>cipal axes of the body.<br />

Spherical top: 1 = 2 = 3<br />

A spherical top is a body hav<strong>in</strong>g three degenerate pr<strong>in</strong>cipal moments of <strong>in</strong>ertia. Such a body has the same<br />

symmetry as the <strong>in</strong>ertia tensor about the center of a uniform sphere. For a sphere it is obvious from the<br />

symmetry that any orientation of three mutually orthogonal axes about the center of the uniform sphere are<br />

equally good pr<strong>in</strong>cipal axes. For a uniform cube the pr<strong>in</strong>cipal axes of the <strong>in</strong>ertia tensor about the center of<br />

mass were shown to be aligned such that they pass through the center of each face, and the three pr<strong>in</strong>cipal<br />

moments are identical; that is, <strong>in</strong>ertially it is equivalent to a spherical top. A less obvious consequence of the<br />

spherical symmetry is that any orientation of three mutually perpendicular axes about the center of mass of<br />

a uniform cube is an equally good pr<strong>in</strong>cipal axis system.


320 CHAPTER 13. RIGID-BODY ROTATION<br />

Symmetric top: 1 = 2 6= 3<br />

The equivalent ellipsoid for a body with two degenerate pr<strong>in</strong>cipal moments of <strong>in</strong>ertia is a spheroid which has<br />

cyl<strong>in</strong>drical symmetry with the cyl<strong>in</strong>drical axis aligned along the third axis. A body with 3 1 = 2 is a<br />

prolate spheroid while a body with 3 1 = 2 is an oblate spheroid. Examples with a prolate spheroidal<br />

equivalent <strong>in</strong>ertial shape are a rugby ball, pencil, or a baseball bat. Examples of an oblate spheroid are an<br />

orange, or a frisbee. A uniform sphere, or a uniform cube, rotat<strong>in</strong>g about a po<strong>in</strong>t displaced from the centerof-mass<br />

also behave <strong>in</strong>ertially like a symmetric top. The cyl<strong>in</strong>drical symmetry of the equivalent spheroid<br />

makes it obvious that any mutually perpendicular axes that are normal to the axis of cyl<strong>in</strong>drical symmetry<br />

areequallygoodpr<strong>in</strong>cipalaxesevenwhenthecrosssection<strong>in</strong>the1−2 plane is square as opposed to circular.<br />

A rotor is a diatomic-molecule shaped body which is a special case of a symmetric top where 1 =0<br />

and 2 = 3 . The rotation of a rotor is perpendicular to the symmetry axis s<strong>in</strong>ce the rotational energy and<br />

angular momentum about the symmetry axis are zero because the pr<strong>in</strong>cipal moment of <strong>in</strong>ertia about the<br />

symmetry axis is zero.<br />

Asymmetric top: 1 6= 2 6= 3<br />

A body where all three pr<strong>in</strong>cipal moments of <strong>in</strong>ertia are dist<strong>in</strong>ct, 1 6= 2 6= 3 is called an asymmetric<br />

top. Some molecules, and nuclei have asymmetric, triaxially-deformed, shapes.<br />

13.10.2 Orthogonality of pr<strong>in</strong>cipal axes<br />

The body-fixed pr<strong>in</strong>cipal axes comprise an orthogonal set, for which the vectors L and ω are simply related.<br />

Components of L and ω can be taken along the three body-fixed axes denoted by Thus for the <br />

pr<strong>in</strong>cipal moment <br />

= (13.46)<br />

Written <strong>in</strong> terms of the <strong>in</strong>ertia tensor<br />

=<br />

3X<br />

= (13.47)<br />

<br />

Similarly the pr<strong>in</strong>cipal moment can be written as<br />

=<br />

3X<br />

= (13.48)<br />

<br />

Multiply the equation 1347 by and sum over gives<br />

X<br />

= X <br />

<br />

(13.49)<br />

Similarly multiply<strong>in</strong>g equation 1348 by and summ<strong>in</strong>g over gives<br />

X<br />

= X (13.50)<br />

<br />

<br />

The left-hand sides of these equations are identical s<strong>in</strong>ce the <strong>in</strong>ertia tensor is symmetric, that is = <br />

Therefore subtract<strong>in</strong>g these equations gives<br />

X<br />

− X =0 (13.51)<br />

<br />

<br />

That is<br />

( − ) X <br />

=0 (13.52)<br />

or<br />

( − ) ω · ω =0 (13.53)


13.11. ANGULAR MOMENTUM L AND ANGULAR VELOCITY ω VECTORS 321<br />

If 6= then<br />

ω · ω =0 (13.54)<br />

which implies that the and pr<strong>in</strong>cipal axes are perpendicular. However, if = then equation<br />

1353 does not require that ω · ω =0, that is, these axes are not necessarily perpendicular, but, with<br />

no loss of generality, these two axes can be chosen to be perpendicular with any orientation <strong>in</strong> the plane<br />

perpendicular to the symmetry axis.<br />

Summariz<strong>in</strong>g the above discussion, the <strong>in</strong>ertia tensor has the follow<strong>in</strong>g properties.<br />

1) Diagonalization may be accomplished by an appropriate rotation of the axes <strong>in</strong> the body.<br />

2) The pr<strong>in</strong>cipal moments (eigenvalues) and pr<strong>in</strong>cipal axes (eigenvectors) are obta<strong>in</strong>ed as roots of the<br />

secular determ<strong>in</strong>ant and are real.<br />

3) The pr<strong>in</strong>cipal axes (eigenvectors) are real and orthogonal.<br />

4) For a symmetric top with two identical pr<strong>in</strong>cipal moments of <strong>in</strong>ertia, any orientation of two orthogonal<br />

axes perpendicular to the symmetry axis are satisfactory eigenvectors.<br />

5) For a spherical top with three identical pr<strong>in</strong>cipal moment of <strong>in</strong>ertia, the pr<strong>in</strong>cipal axes system can<br />

have any orientation with respect to the orig<strong>in</strong>.<br />

13.11 Angular momentum L and angular velocity ω vectors<br />

The angular momentum is a primary observable for rotation. As discussed <strong>in</strong> chapter 135, the angular<br />

momentum L is compactly and elegantly written <strong>in</strong> matrix form us<strong>in</strong>g the tensor algebra relation<br />

⎛<br />

L= ⎝ ⎞ ⎛<br />

11 12 13<br />

21 22 23<br />

⎠ · ⎝ ⎞<br />

1<br />

2<br />

⎠ = {I}·ω (13.55)<br />

31 32 33 3<br />

where ω is the angular velocity, {I} the <strong>in</strong>ertia tensor, and L the correspond<strong>in</strong>g angular momentum.<br />

Two important consequences of equation 1355 are that:<br />

• The angular momentum L and angular velocity ω are not necessarily col<strong>in</strong>ear.<br />

• In general the Pr<strong>in</strong>cipal axis system of the rotat<strong>in</strong>g rigidbodyisnotalignedwitheithertheangular<br />

momentum or angular velocity vectors.<br />

An exception to these statements occurs when the angular velocity ω is aligned along a pr<strong>in</strong>cipal axes<br />

for which the <strong>in</strong>ertia tensor is diagonal, i.e. = , and then both L and ω po<strong>in</strong>t along this pr<strong>in</strong>cipal<br />

axis. In general the angular momentum L and angular velocity ω precess around each other. An important<br />

special case is for torque-free systems where Noether’s theorem implies that the angular momentum vector<br />

L is conserved both <strong>in</strong> magnitude and amplitude. In this case, the angular velocity ω and the Pr<strong>in</strong>cipal axis<br />

system, both precesses around the angular momentum vector L. That is, the body appears to tumble with<br />

respect to the laboratory fixed frame. Understand<strong>in</strong>g rigid-body rotation requires care not to confuse the<br />

body-fixed Pr<strong>in</strong>cipal axis coord<strong>in</strong>ate frame, used to determ<strong>in</strong>e the <strong>in</strong>ertia tensor, and the fixed laboratory<br />

frame where the motion is observed.<br />

13.5 Example: Rotation about the center of mass of a solid cube<br />

It is illustrative to use the <strong>in</strong>ertia tensors of a uniform cube to compute the angular momentum for any<br />

applied angular velocity vector us<strong>in</strong>g equation (1355). If the angular velocity is along the axis, then<br />

us<strong>in</strong>g the <strong>in</strong>ertia tensor for a solid cube, derived earlier, <strong>in</strong> equation (1355) gives the angular momentum to<br />

be<br />

⎛<br />

L = {I}·ω = 1 6 2 ⎝ 1 0 0 ⎞ ⎛<br />

0 1 0 ⎠ · ⎝ 1 ⎞ ⎛<br />

0 ⎠ = 1<br />

0 0 1 0<br />

6 2 ⎝ 1 ⎞<br />

0 ⎠<br />

0<br />

This shows that L and ω are col<strong>in</strong>ear and thus the axis is a pr<strong>in</strong>cipal axis. By symmetry, the and <br />

body fixed axis also must be pr<strong>in</strong>cipal axes.


322 CHAPTER 13. RIGID-BODY ROTATION<br />

Consider that the body is rotated about a diagonal of the cube for which ⎛ the center of mass will be on<br />

the rotation axis. Then the angular velocity vector is written as ω = √ 1 ⎝ 1 ⎞<br />

3<br />

1 ⎠ where the components of<br />

1<br />

q<br />

= = = √ 1 3<br />

with the angular velocity magnitude 2 + 2 + 2 = <br />

⎛<br />

L = {I}·ω = 1 6 2 √ 1 ⎝ 1 0 0 ⎞ ⎛<br />

0 1 0 ⎠ ·<br />

3 0 0 1<br />

⎝ 1 1<br />

1<br />

⎞<br />

⎛<br />

⎠ = 1 6 2 √ 1 3<br />

⎝ 1 1<br />

1<br />

⎞<br />

⎠ = 1 6 2 ω<br />

Note that L and ω aga<strong>in</strong> are col<strong>in</strong>ear show<strong>in</strong>g it also is a pr<strong>in</strong>cipal axis. Moreover, the magnitude of L<br />

is identical for orientations of the rotation axes pass<strong>in</strong>g through the center of mass when centered on<br />

either one face, or the diagonal, of the cube imply<strong>in</strong>g that the pr<strong>in</strong>cipal moments of <strong>in</strong>ertia about these axes<br />

are identical. This illustrates the important property that, when the three pr<strong>in</strong>cipal moments of <strong>in</strong>ertia are<br />

identical, then any orientation of the coord<strong>in</strong>ate system is an equally good pr<strong>in</strong>cipal axis system. That is,<br />

this corresponds to the spherical top where all orientations are pr<strong>in</strong>cipal axes, not just along the obvious<br />

symmetry axes.<br />

13.6 Example: Rotation about the corner of the cube<br />

Let us repeat the above exercise for rotation about one corner of the cube. Consider that the angular<br />

velocity is along the axis. Then example (132) gives the angular momentum to be<br />

⎛<br />

⎞ ⎛<br />

L = {I}·ω = 1 +8 −3 −3<br />

12 2 ⎝ −3 +8 −3 ⎠ · ⎝ 1 ⎞<br />

⎛<br />

0 ⎠ = 1<br />

12 2 ω ⎝ +8 ⎞<br />

−3 ⎠<br />

−3 −3 +8 0<br />

−3<br />

The angular momentum is far from be<strong>in</strong>g aligned with the axis that is, it is not a pr<strong>in</strong>cipal axis.<br />

Consider that the body is rotated with the angular velocity aligned along a⎛diagonal of the cube through<br />

the center of mass on this axis. Then the angular velocity is written as ω = √ 1 ⎝ 1 ⎞<br />

3<br />

1 ⎠ where the components<br />

1<br />

q<br />

of = = = √ 1 3<br />

ensur<strong>in</strong>g that the magnitude equals 2 + 2 + 2 = <br />

⎛<br />

L = {I}·ω = 1<br />

12 2 √ 1 ⎝<br />

3<br />

+8 −3 −3<br />

−3 +8 −3<br />

−3 −3 +8<br />

⎞ ⎛<br />

⎠ ·<br />

⎝ 1 1<br />

1<br />

⎞<br />

⎛<br />

⎠ = 1 12 2 √ 1 3<br />

⎝ 2 2<br />

2<br />

⎞<br />

⎠ = 1 6 2 ω<br />

This is a pr<strong>in</strong>cipal axis s<strong>in</strong>ce L and aga<strong>in</strong> are col<strong>in</strong>ear and the angular momentum is the same as for any<br />

axis through the center of mass of a uniform solid cube due to the high symmetry of the cube. If the angular<br />

velocity is perpendicular to the diagonal of the cube, then, for either of these perpendicular axes, the relation<br />

between and is given by<br />

⎛<br />

L = 1<br />

12 2 √ 1 ⎝<br />

2<br />

+8 −3 −3<br />

−3 +8 −3<br />

−3 −3 +8<br />

⎞ ⎛<br />

⎠ ·<br />

⎝ −1<br />

+1<br />

0<br />

⎞<br />

⎛<br />

⎠ = 1 12 2 √ 1 ⎝ −11<br />

+11<br />

2 0<br />

⎞<br />

⎛<br />

⎠ = 11<br />

12 2 <br />

⎝ −1<br />

+1<br />

0<br />

Note that this must be a pr<strong>in</strong>cipal axis for rotation about a corner of the cube s<strong>in</strong>ce L and ω are col<strong>in</strong>ear.<br />

The angular momentum is the same for both possible orientations of that are perpendicular to the diagonal<br />

through the center of mass. Diagonaliz<strong>in</strong>g the <strong>in</strong>ertia tensor <strong>in</strong> example 132 also gave the above result with<br />

the symmetry axis along the diagonal of the cube.<br />

This example illustrates that it is not necessary to diagonalize the <strong>in</strong>ertia tensor matrix to obta<strong>in</strong> the<br />

pr<strong>in</strong>cipal axes. The corner of the cube has three mutually perpendicular pr<strong>in</strong>cipal axes <strong>in</strong>dependent of the<br />

choice of a body-fixed coord<strong>in</strong>ate frame. The advantage of the pr<strong>in</strong>cipal axis coord<strong>in</strong>ate frame is that the<br />

<strong>in</strong>ertia tensor is diagonal mak<strong>in</strong>g evaluation of the angular momentum trivial. That is, there is no physics<br />

associated with the orientation chosen for the body-fixed coord<strong>in</strong>ate frame, this frame only determ<strong>in</strong>es the<br />

ratio of the components of the <strong>in</strong>ertia tensor along the chosen coord<strong>in</strong>ates. Note that, if a body has an obvious<br />

symmetry, then <strong>in</strong>tuition is a powerful way to identify the pr<strong>in</strong>cipal axis frame.<br />

⎞<br />


13.12. KINETIC ENERGY OF ROTATING RIGID BODY 323<br />

13.12 K<strong>in</strong>etic energy of rotat<strong>in</strong>g rigid body<br />

An important observable is the k<strong>in</strong>etic energy of rotation of a rigid body. Consider a rigid body composed<br />

of particles of mass where =1 2 3 If the body rotates with an <strong>in</strong>stantaneous angular velocity<br />

ω about some fixed po<strong>in</strong>t, with respect to the body coord<strong>in</strong>ate system, and this po<strong>in</strong>t has an <strong>in</strong>stantaneous<br />

translational velocity V with respect to the fixed (<strong>in</strong>ertial) coord<strong>in</strong>ate system, see figure 131, then the<br />

<strong>in</strong>stantaneous velocity v of the particle <strong>in</strong> the fixed frame of reference is given by<br />

v = V + v 00 + ω × r 0 (13.56)<br />

However, for a rigid body, the velocity of a body-fixed po<strong>in</strong>t with respect to the body is zero, that is v 00 =0<br />

thus<br />

v = V + ω × r 0 (13.57)<br />

The total k<strong>in</strong>etic energy is given by<br />

=<br />

= 1 2<br />

X<br />

<br />

1<br />

2 v · v =<br />

<br />

X<br />

X<br />

X<br />

2 + V · ω × r 0 + 1 2<br />

<br />

<br />

1<br />

2 (V + ω × r 0 ) · (V + ω × r 0 )<br />

X<br />

(ω × r 0 ) · (ω × r 0 ) (13.58)<br />

This is a general expression for the k<strong>in</strong>etic energy that is valid for any choice of the orig<strong>in</strong> from which the<br />

body-fixed vectors r 0 are measured. However, if the orig<strong>in</strong> is chosen to be the center of mass, then, and only<br />

then, the middle term cancels. That is, s<strong>in</strong>ce V · ω is <strong>in</strong>dependent of the specific particle, then<br />

Ã<br />

X<br />

<br />

!<br />

X<br />

V · ω × r 0 = V · ω × r 0 <br />

<br />

<br />

<br />

(13.59)<br />

But the def<strong>in</strong>ition of the center of mass is<br />

X<br />

r 0 = R (13.60)<br />

<br />

and R =0<strong>in</strong> the body-fixed frame if the selected po<strong>in</strong>t <strong>in</strong> the body is the center of mass. Thus, when us<strong>in</strong>g<br />

the center of mass frame, the middle term of equation 1358 is zero. Therefore, for the center of mass frame,<br />

the k<strong>in</strong>etic energy separates <strong>in</strong>to two terms <strong>in</strong> the body-fixed frame<br />

where<br />

= + (13.61)<br />

= 1 2<br />

= 1 2<br />

X<br />

2 (13.62)<br />

<br />

X<br />

(ω × r 0 ) · (ω × r 0 )<br />

<br />

The vector identity<br />

(A × B) · (A × B) = 2 2 − (A · B) 2 (13.63)<br />

canbeusedtosimplify <br />

= 1 2<br />

X<br />

<br />

h 2 02 − (ω · r 0 ) 2i (13.64)<br />

<br />

The rotational k<strong>in</strong>etic energy can be expressed <strong>in</strong> terms of components of ω and r 0 <strong>in</strong> the body-fixed<br />

frame. Also the follow<strong>in</strong>g formulae are greatly simplified if r 0 =( ) <strong>in</strong> the rotat<strong>in</strong>g body-fixed frame


324 CHAPTER 13. RIGID-BODY ROTATION<br />

is written <strong>in</strong> the form r 0 =( 1 2 3 ) where the axes are def<strong>in</strong>ed by the numbers 1 2 3 rather than<br />

. In this notation the rotational k<strong>in</strong>etic energy is written as<br />

⎡Ã !Ã ! Ã !<br />

= 1 X<br />

⎛ ⎞⎤<br />

X X X<br />

<br />

⎣ 2 2 − <br />

⎝ X <br />

⎠⎦ (13.65)<br />

2<br />

<br />

<br />

<br />

<br />

<br />

Assume the Kronecker delta relation<br />

=<br />

3X<br />

(13.66)<br />

<br />

where =1 if = and =0 if 6= <br />

Then the k<strong>in</strong>etic energy can be written more compactly<br />

⎡Ã !Ã ! Ã !<br />

= 1 X<br />

⎛ ⎞⎤<br />

X X X<br />

<br />

⎣ 2 2 − <br />

⎝ X <br />

⎠⎦<br />

2<br />

<br />

<br />

<br />

<br />

<br />

Ã<br />

= 1 X 3X<br />

3X<br />

!<br />

#<br />

<br />

"( ) 2 − ( )( )<br />

2<br />

<br />

<br />

"<br />

= 1 3X <br />

" Ã<br />

X<br />

3X<br />

! ##<br />

2 − (13.67)<br />

2<br />

<br />

<br />

<br />

The term <strong>in</strong> the outer square brackets is the <strong>in</strong>ertia tensor def<strong>in</strong>ed <strong>in</strong> equation 1312 for a discrete body. The<br />

<strong>in</strong>ertia tensor components for a cont<strong>in</strong>uous body are given by equation 1313.<br />

Thus the rotational component of the k<strong>in</strong>etic energy can be written <strong>in</strong> terms of the <strong>in</strong>ertia tensor as<br />

= 1 2<br />

3X<br />

(13.68)<br />

Note that when the <strong>in</strong>ertia tensor is diagonal ,then the evaluation of the k<strong>in</strong>etic energy simplifies to<br />

= 1 2<br />

<br />

3X<br />

2 (13.69)<br />

which is the familiar relation <strong>in</strong> terms of the scalar moment of <strong>in</strong>ertia discussed <strong>in</strong> elementary mechanics.<br />

Equation 1368 also can be factored <strong>in</strong> terms of the angular momentum L.<br />

= 1 X<br />

= 1 X X<br />

= 1 X<br />

(13.70)<br />

2<br />

2<br />

2<br />

<br />

<br />

As mentioned earlier, tensor algebra is an elegant and compact way of express<strong>in</strong>g such matrix operations.<br />

Thus it is possible to express the rotational k<strong>in</strong>etic energy as<br />

⎛<br />

= 1 ¡ ¢<br />

1 2 3 · ⎝ ⎞ ⎛<br />

11 12 13<br />

21 22 23<br />

⎠ · ⎝ ⎞<br />

1<br />

2<br />

⎠ (13.71)<br />

2<br />

31 32 33 3<br />

≡ T = 1 ω ·{I}·ω (13.72)<br />

2<br />

where the rotational energy T is a scalar. Us<strong>in</strong>g equation 1355 the rotational component of the k<strong>in</strong>etic<br />

energy also can be written as<br />

≡ T = 1 2 ω · L (13.73)<br />

which is the same as given by (1370). It is <strong>in</strong>terest<strong>in</strong>g to realize that even though L = {I}·ω is the <strong>in</strong>ner<br />

product of a tensor and a vector, it is a vector as illustrated by the fact that the <strong>in</strong>ner product = 1 2 ω·L =<br />

1<br />

2 ω · ({I}·ω) is a scalar. Note that the translational k<strong>in</strong>etic energy must be added to the rotational<br />

k<strong>in</strong>etic energy to get the total k<strong>in</strong>etic energy as given by equation 1361


13.13. EULER ANGLES 325<br />

13.13 Euler angles<br />

The description of rigid-body rotation is greatly facilitated<br />

by transform<strong>in</strong>g from the space-fixed coord<strong>in</strong>ate<br />

frame (ˆx ŷ ẑ) toarotat<strong>in</strong>gbody-fixed coord<strong>in</strong>ate frame<br />

¡ˆ1 ˆ2 ˆ3 ¢ for which the <strong>in</strong>ertia tensor is diagonal. Appendix<br />

<strong>in</strong>troduced the rotation matrix {λ} which can be<br />

used to rotate between the space-fixed coord<strong>in</strong>ate system,<br />

which is stationary, and the <strong>in</strong>stantaneous bodyfixed<br />

frame which is rotat<strong>in</strong>g with respect to the spacefixed<br />

frame. The transformation can be represented by<br />

a matrix equation<br />

¡ˆ1 ˆ2 ˆ3 ¢ = {λ}·(ˆx ŷ ẑ) (13.74)<br />

where the space-fixed system is identified by unit vectors<br />

(ˆx ŷ ẑ) while ¡ˆ1 ˆ2 ˆ3 ¢ def<strong>in</strong>es unit vectors <strong>in</strong> the rotated<br />

body-fixed system. The rotation matrix {λ} completely<br />

describes the <strong>in</strong>stantaneous relative orientation of the<br />

two systems. Rigid-body rotation requires three <strong>in</strong>dependent<br />

angular parameters that specify the orientation<br />

of the rigid body such that the correspond<strong>in</strong>g orthogonal<br />

transformation matrix is proper, that is, it has a<br />

determ<strong>in</strong>ant || =+1as given by equation (33).<br />

As discussed <strong>in</strong> Appendix 2, the9 component rotation<br />

matrix <strong>in</strong>volves only three <strong>in</strong>dependent angles.<br />

There are many possible choices for these three angles.<br />

It is convenient to use the Euler angles, (also<br />

called Eulerian angles) shown <strong>in</strong> figure 133. 1 The Euler<br />

angles are generated by a series of three rotations that<br />

rotate from the space-fixed (ˆx ŷ ẑ) system to the bodyfixed<br />

¡ˆ1 ˆ2 ˆ3 ¢ system. The rotation must be such that<br />

the space-fixed axis rotates by an angle to align with<br />

the body-fixed 3 axis. This can be performed by rotat<strong>in</strong>g<br />

through an angle about the ˆn ≡ ẑ × ˆ3 direction, where<br />

ẑ and ˆ3 designate the unit vectors along the “” axes<br />

Figure 13.3: The − − sequence of rotations<br />

correspond<strong>in</strong>g to the Eulerian angles<br />

( ). The first rotation about the spacefixed<br />

z axis (blue) is from the -axis (blue) to the<br />

l<strong>in</strong>e of nodes n (green). The second rotation <br />

about the l<strong>in</strong>e of nodes (green) is from the spacefixed<br />

axis (blue) to the body-fixed 3-axis (red).<br />

The third rotation about the body-fixed 3-axis<br />

(red) is from the l<strong>in</strong>e of nodes (green) to the bodyfixed<br />

1 axis (red).<br />

of the space and body fixed frames respectively. The<br />

unit vector ˆn ≡ ẑ × ˆ3 is the vector normal to the plane<br />

def<strong>in</strong>ed by the ẑ and ˆ3 unit vectors and this unit vector ˆn = ẑ × ˆ3 is called the l<strong>in</strong>e of nodes. The chosen<br />

convention is that the unit vector ˆn = ẑ × ˆ3 is along the “” axis of an <strong>in</strong>termediate-axis frame designated<br />

by ¡ˆn ŷ 0 ẑ ¢ , that is, the unit vector ˆn = ẑ × ˆ3 plus the unit vectors ŷ 0 and ẑ are <strong>in</strong> the same plane as the ẑ<br />

and ˆ3 unit vectors. The sequence of three rotations is performed as summarized below.<br />

1) Rotation about the space-fixed ẑ axis from the space ˆx axis to the l<strong>in</strong>e of nodes ˆn : The<br />

first rotation (x y z) · λ <br />

→ (n y 0 z) is <strong>in</strong> a right-handed direction through an angle about the space-fixed<br />

z axis. S<strong>in</strong>ce the rotation takes place <strong>in</strong> the x − y plane, the transformation matrix is<br />

⎛<br />

{λ } = ⎝<br />

cos s<strong>in</strong> 0<br />

− s<strong>in</strong> cos 0<br />

0 0 1<br />

⎞<br />

⎠ (13.75)<br />

1 The space-fixed coord<strong>in</strong>ate frame and the body-fixed coord<strong>in</strong>ate frames are unambiguously def<strong>in</strong>ed, that is, the space-fixed<br />

frame is stationary while the body-fixed frame is the pr<strong>in</strong>cipal-axis frame of the body. There are several possible <strong>in</strong>termediate<br />

frames that can be used to def<strong>in</strong>e the Euler angles. The − − sequence of rotations, used here, is used <strong>in</strong> most physics<br />

textbooks <strong>in</strong> classical mechanics. Unfortunately scientists and eng<strong>in</strong>eers use slightly different conventions for def<strong>in</strong><strong>in</strong>g the Euler<br />

angles. As discussed <strong>in</strong> Appendix A of "<strong>Classical</strong> <strong>Mechanics</strong>" by Goldste<strong>in</strong>, nuclear and particle physicists have adopted the<br />

− − sequence of rotations while the US and UK aerodynamicists have adopted a − − sequence of rotations.


326 CHAPTER 13. RIGID-BODY ROTATION<br />

This leads to the <strong>in</strong>termediate coord<strong>in</strong>ate system (n y 0 z) where the rotated x axis now is col<strong>in</strong>ear with the<br />

n axis of the <strong>in</strong>termediate frame, that is, the l<strong>in</strong>e of nodes.<br />

(n y 0 z) ={λ }·(x y z) (13.76)<br />

The precession angular velocity ˙ istherateofchangeofangleofthel<strong>in</strong>eofnodeswithrespecttothespace<br />

axis about the space-fixed axis.<br />

2) Rotation about the l<strong>in</strong>e of nodes ˆn from the space ẑ axis to the body-fixed ˆ3 axis: The<br />

second rotation<br />

(n y 0 z) · → (n y 00 3) (13.77)<br />

is <strong>in</strong> a right-handed direction through the angle about the ˆn axis (l<strong>in</strong>e of nodes) so that the “” axis becomes<br />

col<strong>in</strong>ear with the body-fixed ˆ3 axis. Because the rotation now is <strong>in</strong> the ẑ−ˆ3 plane, the transformation matrix<br />

is<br />

⎛<br />

{λ } = ⎝ 1 0 0 ⎞<br />

0 cos s<strong>in</strong> ⎠ (13.78)<br />

0 − s<strong>in</strong> cos <br />

The l<strong>in</strong>e of nodes which is at the <strong>in</strong>tersection of the space-fixed and body-fixed planes, shown <strong>in</strong> figure 133<br />

po<strong>in</strong>ts <strong>in</strong> the ˆn = ẑ × ˆ3 direction. The new “” axis now is the body-fixed ˆ3 axis. The angular velocity ˙ is<br />

therateofchangeofangleof thebody-fixed ˆ3-axis relative to the space-fixed ẑ-axis about the l<strong>in</strong>e of nodes.<br />

3) Rotation about the body-fixed ˆ3 axis from the l<strong>in</strong>e of nodes to the body-fixed ˆ1 axis: The<br />

third rotation<br />

(n y 00 3) · → (ˆ1 ˆ2 ˆ3) (13.79)<br />

is <strong>in</strong> a right-handed direction through the angle about the new body-fixed ˆ3 axis This third rotation<br />

transforms the rotated <strong>in</strong>termediate (n y 00 3) frame to f<strong>in</strong>al body-fixed coord<strong>in</strong>ate system (ˆ1 ˆ2 ˆ3) The<br />

transformation matrix is<br />

⎛<br />

{λ } = ⎝<br />

cos s<strong>in</strong> 0<br />

− s<strong>in</strong> cos 0<br />

0 0 1<br />

⎞<br />

⎠ (13.80)<br />

The sp<strong>in</strong> angular velocity ˙ is the rate of change of the angle of the body-fixed 1-axis with respect to the<br />

l<strong>in</strong>e of nodes about the body-fixed 3 axis.<br />

The total rotation matrix {λ} is given by<br />

{λ} = {λ }·{λ }·{λ } (13.81)<br />

Thus the complete rotation from the space-fixed (x y z) axis system to the body-fixed (1 2 3) axis system<br />

is given by<br />

(1 2 3) ={λ}·(x y z) (13.82)<br />

where {λ} is given by the triple product equation (1381) lead<strong>in</strong>gtotherotationmatrix<br />

⎛<br />

{λ} = ⎝<br />

cos cos − s<strong>in</strong> cos s<strong>in</strong> s<strong>in</strong> cos +cos cos s<strong>in</strong> s<strong>in</strong> s<strong>in</strong> <br />

− cos s<strong>in</strong> − s<strong>in</strong> cos cos − s<strong>in</strong> s<strong>in</strong> +cos cos cos s<strong>in</strong> cos <br />

s<strong>in</strong> s<strong>in</strong> − cos s<strong>in</strong> cos <br />

⎞<br />

⎠ (13.83)<br />

The <strong>in</strong>verse transformation from the body-fixed axis system to the space-fixed axis system is given by<br />

(x y z) ={λ} −1 · (1 2 3) (13.84)<br />

wherethe<strong>in</strong>versematrix{λ} −1 equals the transposed rotation matrix {λ} ,thatis,<br />

⎛<br />

{λ} −1 = {λ} = ⎝<br />

cos cos − s<strong>in</strong> cos s<strong>in</strong> − cos s<strong>in</strong> − s<strong>in</strong> cos cos s<strong>in</strong> s<strong>in</strong> <br />

s<strong>in</strong> cos +cos cos s<strong>in</strong> − s<strong>in</strong> s<strong>in</strong> +cos cos cos − cos s<strong>in</strong> <br />

s<strong>in</strong> s<strong>in</strong> s<strong>in</strong> cos cos <br />

⎞<br />

⎠ (13.85)


13.14. ANGULAR VELOCITY ω 327<br />

Tak<strong>in</strong>g the product {λ}{λ} −1 =1shows that the rotation matrix is a proper, orthogonal, unit matrix.<br />

The use of three different coord<strong>in</strong>ate systems, space-fixed, the <strong>in</strong>termediate l<strong>in</strong>e of nodes, and the bodyfixed<br />

frame can be confus<strong>in</strong>g at first glance. Basically the angle specifies the rotation about the space-fixed<br />

axis between the space-fixed axis and the l<strong>in</strong>e of nodes of the Euler angle <strong>in</strong>termediate frame. The angle<br />

specifies the rotation about the body-fixed 3 axis between the l<strong>in</strong>e of nodes and the body-fixed 1 axis. Note<br />

that although the space-fixed and body-fixed axes systems each are orthogonal, the Euler angle basis <strong>in</strong><br />

general is not orthogonal. For rigid-body rotation the rotation angle about the space-fixed axis is time<br />

dependent, that is, the l<strong>in</strong>e of nodes is rotat<strong>in</strong>g with an angular velocity ˙ with respect to the space-fixed<br />

coord<strong>in</strong>ate frame. Similarly the body-fixed coord<strong>in</strong>ate frame is rotat<strong>in</strong>g about the body-fixed 3 axis with<br />

angular velocity ˙ relative to the l<strong>in</strong>e of nodes.<br />

13.7 Example: Euler angle transformation<br />

The def<strong>in</strong>ition of the Euler angles can be confus<strong>in</strong>g, therefore it is useful to illustrate their use for a<br />

rotational transformation of a primed frame ( 0 0 0 ) to an unprimed frame ( ) Assume the first<br />

rotation about the 0 axis, is =30 ◦ ⎛<br />

⎞<br />

=<br />

⎜<br />

⎝<br />

√<br />

3<br />

2<br />

− 1 2<br />

1<br />

2<br />

0<br />

√<br />

3<br />

2<br />

0<br />

0 0 1<br />

Let the second rotation be =45 ◦ about the l<strong>in</strong>e of nodes, that is, the <strong>in</strong>termediate ” axis. Then<br />

⎟<br />

⎠<br />

⎞<br />

⎛<br />

1 0 0<br />

0 − √ 1 2<br />

√2 1<br />

= ⎝ 0 √2 1 √2 1<br />

⎠<br />

Let the third rotation be =90 ◦ about the axis.<br />

⎛<br />

= ⎝ 0 1 0 ⎞<br />

−1 0 0 ⎠<br />

0 0 1<br />

Thus the net rotation corresponds to = <br />

⎛ √ ⎞<br />

3<br />

⎛<br />

⎞ ⎛<br />

1<br />

⎜<br />

2 2<br />

0 1 0 0<br />

= ⎝ − 1 √<br />

3<br />

⎟<br />

2 2<br />

0 ⎠ ⎝ 0 √2 1 √2 1<br />

⎠ ⎝ 0 1 0 ⎞ ⎛<br />

−1 0 0 ⎠ = ⎝<br />

0 0 1 0 − √ 1 √2 1<br />

2<br />

0 0 1<br />

√<br />

− 4√ 1 2<br />

1<br />

2 3<br />

1<br />

4√<br />

2<br />

− 4√ 1 6 −<br />

1 1<br />

1<br />

2<br />

√ 2<br />

2 0<br />

1<br />

2<br />

⎞<br />

4√<br />

√ 6 ⎠<br />

2<br />

13.14 Angular velocity ω<br />

It is useful to relate the rigid-body equations of motion <strong>in</strong> the space-fixed (ˆx ŷ ẑ) coord<strong>in</strong>ate system to<br />

those <strong>in</strong> the body-fixed (ê 1 ê 2 ê 3 ) coord<strong>in</strong>ate system where the pr<strong>in</strong>cipal axis <strong>in</strong>ertia tensor is def<strong>in</strong>ed. It<br />

was shown <strong>in</strong> appendix that an <strong>in</strong>f<strong>in</strong>itessimal rotation can be represented by a vector. Thus the time<br />

derivatives of these rotation angles can be associated with the components of the angular velocity ω where<br />

the precession = ˙, thenutation = ˙, andthesp<strong>in</strong> = ˙. Unfortunately the coord<strong>in</strong>ates ( )<br />

are with respect to mixed coord<strong>in</strong>ate frames and thus are not orthogonal axes. That is, the Euler angular<br />

velocities are expressed <strong>in</strong> different coord<strong>in</strong>ate frames, where the precession ˙ is around the space-fixed ẑ<br />

axis measured relative to the ˆx-axis, the sp<strong>in</strong> ˙ is around the body-fixed ê 3 axis relative to the rotat<strong>in</strong>g<br />

l<strong>in</strong>e-of-nodes, and the nutation ˙ is the angular velocity between the ẑ and ê 3 axes and po<strong>in</strong>ts along the<br />

<strong>in</strong>stantaneous l<strong>in</strong>e-of-nodes <strong>in</strong> the ê 3 × ẑ direction. By reference to figure 133 it can be seen that the<br />

components along the body-fixed axes are as given <strong>in</strong> Table 131.<br />

Table 131; Euler angular velocity components <strong>in</strong> the body-fixed frame<br />

Precession ˙ Nutation ˙ Sp<strong>in</strong> ˙<br />

˙ 1 = ˙ s<strong>in</strong> s<strong>in</strong> ˙1 = ˙ cos ˙1 =0<br />

˙ 2 = ˙ s<strong>in</strong> cos ˙2 = −˙ s<strong>in</strong> ˙2 =0<br />

˙ 3 = ˙ cos ˙3 =0 ˙3 = ˙


328 CHAPTER 13. RIGID-BODY ROTATION<br />

Note that the precession angular velocity ˙ is the angular velocity that the body-fixed ê 3 and ẑ × ˆ3 axes<br />

precess around the space-fixed ẑ axis. Table 131 gives the Euler angular velocities required to calculate<br />

the components of the angular velocity ω for the body-fixed (1 2 3) axis system. Collect<strong>in</strong>g the <strong>in</strong>dividual<br />

components of ω gives the components of the angular velocity of the body, relative to the space-fixed axes,<br />

<strong>in</strong> the body-fixed axis system (1 2 3)<br />

1 = ˙ 1 + ˙ 1 + ˙ 1 = ˙ s<strong>in</strong> s<strong>in</strong> + ˙ cos (13.86)<br />

2 = ˙ 2 + ˙ 2 + ˙ 2 = ˙ s<strong>in</strong> cos − ˙ s<strong>in</strong> (13.87)<br />

3 = ˙ 3 + ˙ 3 + ˙ 3 = ˙ cos + ˙ (13.88)<br />

Theangularvelocityofthebodyaboutthebody-fixed 3-axis, 3 , is the sum of the projection of the<br />

precession angular velocity of the l<strong>in</strong>e-of-nodes ˙ with respect to the space-fixed x-axis, plus the angular<br />

velocity ˙ of the body-fixed 3-axis with respect to the rotat<strong>in</strong>g l<strong>in</strong>e-of-nodes.<br />

Similarly, the components of the body angular velocity ω for the space-fixed axis system ( ) can be<br />

derived to be<br />

= ˙ cos + ˙ s<strong>in</strong> s<strong>in</strong> (13.89)<br />

= ˙ s<strong>in</strong> − ˙ s<strong>in</strong> cos (13.90)<br />

= ˙ + ˙ cos (13.91)<br />

Note that when =0then the Euler angles are s<strong>in</strong>gular <strong>in</strong> that the space-fixed axis is parallel with<br />

the body-fixed 3 axis and there is no way of dist<strong>in</strong>guish<strong>in</strong>g between precession ˙ and sp<strong>in</strong> ˙, lead<strong>in</strong>gto<br />

= 3 = ˙ + ˙. When = then the axis and 3 axis are antiparallel and = ˙ − ˙ = − 3 . The other<br />

special case is when cos =0for which the Euler angle system is orthogonal and the space-fixed = ˙,<br />

that is, it equals the precession, while the body-fixed 3 = ˙, that is, it equals the sp<strong>in</strong>. When the Euler<br />

angle basis is not orthogonal then equations (1386 − 88) and (1389 − 91) are needed for express<strong>in</strong>g the<br />

Euler equations of motion <strong>in</strong> either the body-fixed frame or the space-fixed frame respectively.<br />

Equations 1386−88 for the components of the angular velocity <strong>in</strong> the body-fixedframecanbeexpressed<br />

<strong>in</strong> terms of the Euler angle velocities <strong>in</strong> a matrix form as<br />

⎛<br />

⎝ ⎞ ⎛<br />

1<br />

2<br />

⎠ = ⎝<br />

3<br />

s<strong>in</strong> s<strong>in</strong> cos 0<br />

s<strong>in</strong> cos − s<strong>in</strong> 0<br />

cos 0 1<br />

⎞ ⎛<br />

⎠ · ⎝<br />

˙<br />

˙<br />

˙<br />

⎞<br />

⎠ (13.92)<br />

aga<strong>in</strong> note that the transformation matrix is not orthogonal which is to be expected s<strong>in</strong>ce the Euler angular<br />

velocities are about axes that do not form a rectangular system of coord<strong>in</strong>ates. Similarly equations 1389−91<br />

for the angular velocity <strong>in</strong> the space-fixed frame can be expressed <strong>in</strong> terms of the Euler angle velocities <strong>in</strong><br />

matrix form as<br />

⎛ ⎞ ⎛<br />

⎞ ⎛ ⎞<br />

⎝ <br />

<br />

<br />

⎠ = ⎝<br />

0 cos s<strong>in</strong> s<strong>in</strong> <br />

0 s<strong>in</strong> s<strong>in</strong> cos <br />

1 0 cos<br />

⎠ · ⎝<br />

˙<br />

˙<br />

˙<br />

⎠ (13.93)<br />

13.15 K<strong>in</strong>etic energy <strong>in</strong> terms of Euler angular velocities<br />

The k<strong>in</strong>etic energy is a scalar quantity and thus is the same <strong>in</strong> both stationary and rotat<strong>in</strong>g frames of<br />

reference. It is much easier to evaluate the k<strong>in</strong>etic energy <strong>in</strong> the rotat<strong>in</strong>g Pr<strong>in</strong>cipal-axis frame s<strong>in</strong>ce the<br />

<strong>in</strong>ertia tensor is diagonal <strong>in</strong> the Pr<strong>in</strong>cipal-axis frame as given <strong>in</strong> equation 1369<br />

= 1 2<br />

3X<br />

2 (13.94)<br />

Us<strong>in</strong>g equation 1386 − 88 for the body-fixed angular velocities gives the rotational k<strong>in</strong>etic energy <strong>in</strong> terms<br />

of the Euler angular velocities and pr<strong>in</strong>cipal-frame moments of <strong>in</strong>ertia to be<br />

= 1 ∙ ³ ³ ³ ´2¸<br />

1 ˙ s<strong>in</strong> s<strong>in</strong> + ˙ cos ´2<br />

+ 2 ˙ s<strong>in</strong> cos − ˙ s<strong>in</strong> ´2<br />

+ 3 ˙ cos + ˙ (13.95)<br />

2


13.16. ROTATIONAL INVARIANTS 329<br />

13.16 Rotational <strong>in</strong>variants<br />

The scalar properties of a rotat<strong>in</strong>g body, such as mass Lagrangian , and Hamiltonian are rotationally<br />

<strong>in</strong>variant, that is, they are the same <strong>in</strong> any body-fixed or laboratory-fixed coord<strong>in</strong>ate frame. This fact also<br />

applies to scalar products of all vector observables such as angular momentum. For example the scalar<br />

product<br />

L · L = 2 (13.96)<br />

where is the root mean square value of the angular momentum. An example of a scalar <strong>in</strong>variant is the<br />

scalar product of the angular velocity<br />

ω · ω = 2 (13.97)<br />

where 2 is the mean square angular velocity. The scalar product · = || 2 canbecalculatedus<strong>in</strong>gthe<br />

Euler-angle velocities for the body-fixed frame, equations 1386 − 88, tobe<br />

ω · ω = || 2 = 2 1 + 2 2 + 2 3 = ˙ 2 + ˙ 2 + ˙ 2 +2˙ ˙ cos <br />

Similarly, the scalar product can be calculated us<strong>in</strong>g the Euler angle velocities for the space-fixed frame<br />

us<strong>in</strong>g equations 1389 − 91.<br />

ω · ω = || 2 = 2 + 2 + 2 = ˙ 2 + ˙ 2 + ˙ 2 +2˙ ˙ cos <br />

This shows the obvious result that the scalar product · = || 2 is <strong>in</strong>variant to rotations of the coord<strong>in</strong>ate<br />

frame, that is, it is identical when evaluated <strong>in</strong> either the space-fixed, or body-fixed frames.<br />

Note that for =0,theˆ3 and ˆ axes are parallel, and perpendicular to the ˆ axis, then<br />

³<br />

|| 2 ´2<br />

= ˙ + ˙ + ˙2<br />

For the case when = 180 ◦ ,theˆ3 and ˆ axes are antiparallel, and perpendicular to the ˆ axis, then<br />

³<br />

|| 2 ´2<br />

= ˙ − ˙ + ˙2<br />

For the case when =90 ◦ ,theˆ3 , ˆ, andˆ axes are mutually perpendicular, that is, orthogonal, and then<br />

|| 2 = ˙ 2 + ˙ 2 + ˙ 2<br />

The time-averaged shape of a rapidly-rotat<strong>in</strong>g body, as seen <strong>in</strong> the fixed <strong>in</strong>ertial frame, is very different<br />

from the actual shape of the body, and this difference depends on the rotational frequency. For example, a<br />

pencil rotat<strong>in</strong>g rapidly about an axis perpendicular to the body-fixed symmetry axis has an average shape<br />

that is a flat disk <strong>in</strong> the laboratory frame which bears little resemblance to a pencil. The actual shape of the<br />

pencil could be determ<strong>in</strong>ed by tak<strong>in</strong>g high-speed photographs which display the <strong>in</strong>stantaneous body-fixed<br />

shape of the object at given times. Unfortunately for fast rotation, such as rotation of a molecule or a<br />

nucleus, it is not possible to take photographs with sufficient speed and spatial resolution to observe the<br />

<strong>in</strong>stantaneous shape of the rotat<strong>in</strong>g body. What is measured is the average shape of the body as seen <strong>in</strong> the<br />

fixed laboratory frame. In pr<strong>in</strong>ciple the shape observed <strong>in</strong> the fixed <strong>in</strong>ertial frame can be related to the shape<br />

<strong>in</strong> the body-fixed frame, but this requires know<strong>in</strong>g the body-fixedshapewhich<strong>in</strong>generalisnotknown.For<br />

example, a deformed nucleus may be both vibrat<strong>in</strong>g and rotat<strong>in</strong>g about some triaxially deformed average<br />

shape which is a function of the rotational frequency. This is not apparent from the shapes measured <strong>in</strong> the<br />

fixed frame for each of the excited states.<br />

The fact that scalar products are rotationally <strong>in</strong>variant, provides a powerful means of transform<strong>in</strong>g products<br />

of observables <strong>in</strong> the body-fixed frame, to those <strong>in</strong> the laboratory frame. In 1971 Cl<strong>in</strong>e developed<br />

a powerful model-<strong>in</strong>dependent method that utilizes rotationally-<strong>in</strong>variant products of the electromagnetic<br />

quadrupole operator 2 to relate the electromagnetic 2 properties for the observed levels of a rotat<strong>in</strong>g<br />

nucleus measured <strong>in</strong> the laboratory frame, to the electromagnetic 2 properties of the deformed rotat<strong>in</strong>g<br />

nucleus measured <strong>in</strong> the body-fixed frame.[Cli71, Cli72, Cli86] The method uses the fact that scalar products<br />

of the electromagnetic multipole operators are rotationally <strong>in</strong>variant. This allows transform<strong>in</strong>g scalar products<br />

of a complete set of measured electromagnetic matrix elements, measured <strong>in</strong> the laboratory frame, <strong>in</strong>to


330 CHAPTER 13. RIGID-BODY ROTATION<br />

the electromagnetic properties <strong>in</strong> the body-fixed frame of the rotat<strong>in</strong>g nucleus. These rotational <strong>in</strong>variants<br />

provide a model-<strong>in</strong>dependent determ<strong>in</strong>ation of the magnitude, triaxiality, and vibrational amplitudes of the<br />

average shapes <strong>in</strong> the body-fixed frame for <strong>in</strong>dividual observed nuclear states that may be undergo<strong>in</strong>g both<br />

rotation and vibration. When the bombard<strong>in</strong>g energy is below the Coulomb barrier, the scatter<strong>in</strong>g of a<br />

projectile nucleus by a target nucleus is due purely to the electromagnetic <strong>in</strong>teraction s<strong>in</strong>ce the distance<br />

of closest approach exceeds the range of the nuclear force. For such pure Coulomb collisions, the electromagnetic<br />

excitation of collective nuclei populates many excited states, as illustrated <strong>in</strong> figure 1413, with<br />

cross sections that are a direct measure of the 2 matrix elements. These measured matrix elements are<br />

precisely those required to evaluate, <strong>in</strong> the laboratory frame, the 2 rotational <strong>in</strong>variants from which it is<br />

possible to deduce the <strong>in</strong>tr<strong>in</strong>sic quadrupole shapes of the rotat<strong>in</strong>g-vibrat<strong>in</strong>g nuclear states <strong>in</strong> the body-fixed<br />

frame[Cli86].<br />

13.17 Euler’s equations of motion for rigid-body rotation<br />

Rigid-body rotation can be confus<strong>in</strong>g <strong>in</strong> that two coord<strong>in</strong>ate frames are <strong>in</strong>volved and, <strong>in</strong> general, the angular<br />

velocity and angular momentum are not aligned. The motion of the rigid body is observed <strong>in</strong> the space-fixed<br />

<strong>in</strong>ertial frame whereas it is simpler to calculate the equations of motion <strong>in</strong> the body-fixed pr<strong>in</strong>cipal axis<br />

frame, for which the <strong>in</strong>ertia tensor is known and is constant. The rigid body is rotat<strong>in</strong>g about the angular<br />

velocity vector ω, which is not aligned with the angular momentum L. For torque-free motion, L is conserved<br />

and has a fixed orientation <strong>in</strong> the space-fixed axis system. Euler’s equations of motion, presented below,<br />

are given <strong>in</strong> the body-fixed frame for which the <strong>in</strong>ertial tensor is known s<strong>in</strong>ce this simplifies solution of the<br />

equations of motion. However, this solution has to be rotated back <strong>in</strong>to the space-fixed frame to describe<br />

the rotational motion as seen by an observer <strong>in</strong> the <strong>in</strong>ertial frame.<br />

This chapter has <strong>in</strong>troduced the <strong>in</strong>ertial properties of a rigid body, as well as the Euler angles for<br />

transform<strong>in</strong>g between the body-fixed and <strong>in</strong>ertial frames of reference. This has prepared the stage for<br />

solv<strong>in</strong>g the equations of motion for rigid-body motion, namely, the dynamics of rotational motion about a<br />

body-fixed po<strong>in</strong>t under the action of external forces. The Euler angles are used to specify the <strong>in</strong>stantaneous<br />

orientation of the rigid body.<br />

In Newtonian mechanics, the rotational motion is governed by the equivalent Newton’s second law given<br />

<strong>in</strong> terms of the external torque N and angular momentum L<br />

µ L<br />

N =<br />

(13.98)<br />

<br />

Note that this relation is expressed <strong>in</strong> the <strong>in</strong>ertial space-fixed frame of reference, not the non-<strong>in</strong>ertial bodyfixed<br />

frame. The subscript is added to emphasize that this equation is written <strong>in</strong> the <strong>in</strong>ertial space-fixed<br />

frame of reference. However, as already discussed, it is much more convenient to transform from the spacefixed<br />

<strong>in</strong>ertial frame to the body-fixed frame for which the <strong>in</strong>ertia tensor of the rigid body is known. Thus the<br />

next stage is to express the rotational motion <strong>in</strong> terms of the body-fixed frame of reference. For simplicity,<br />

translational motion will be ignored.<br />

The rate of change of angular momentum can be written <strong>in</strong> terms of the body-fixed value, us<strong>in</strong>g the<br />

transformation from the space-fixed <strong>in</strong>ertial frame (ˆx ŷ ẑ) to the rotat<strong>in</strong>g frame (ê 1 ê 2 ê 3 ) as given <strong>in</strong><br />

chapter 103,<br />

N =<br />

µ L<br />

=<br />

<br />

<br />

<br />

However, the body axis ê is chosen to be the pr<strong>in</strong>cipal axis such that<br />

µ L<br />

+ ω × L (13.99)<br />

<br />

<br />

= (13.100)<br />

where the pr<strong>in</strong>cipal moments of <strong>in</strong>ertia are written as . Thus the equation of motion can be written us<strong>in</strong>g<br />

the body-fixed coord<strong>in</strong>ate system as<br />

ê 1 ê 2 ê 3<br />

N = 1 ˙ 1 ê 1 + 2 ˙ 2 ê 2 + 3 ˙ 3 ê 3 +<br />

1 2 3<br />

(13.101)<br />

¯ 1 1 2 2 3 3<br />

¯¯¯¯¯¯<br />

= ( 1 ˙ 1 − ( 2 − 3 ) 2 3 ) ê 1 +( 2 ˙ 2 − ( 3 − 1 ) 3 1 ) ê 2 +( 3 ˙ 3 − ( 1 − 2 ) 1 2 ) ê 3 (13.102)


13.18. LAGRANGE EQUATIONS OF MOTION FOR RIGID-BODY ROTATION 331<br />

where the components <strong>in</strong> the body-fixed axes are given by<br />

1 = 1 ˙ 1 − ( 2 − 3 ) 2 3 (13.103)<br />

2 = 2 ˙ 2 − ( 3 − 1 ) 3 1<br />

3 = 3 ˙ 3 − ( 1 − 2 ) 1 2<br />

These are the Euler equations for rigid body <strong>in</strong> a force field expressed <strong>in</strong> the body-fixed coord<strong>in</strong>ate<br />

frame. They are applicable for any applied external torque N.<br />

The motion of a rigid body depends on the structure of the body only via the three pr<strong>in</strong>cipal moments<br />

of <strong>in</strong>ertia 1 2 and 3 Thus all bodies hav<strong>in</strong>g the same pr<strong>in</strong>cipal moments of <strong>in</strong>ertia will behave exactly the<br />

same even though the bodies may have very different shapes. As discussed earlier, the simplest geometrical<br />

shape of a body hav<strong>in</strong>g three different pr<strong>in</strong>cipal moments is a homogeneous ellipsoid. Thus, the rigid-body<br />

motion often is described <strong>in</strong> terms of the equivalent ellipsoid that has the same pr<strong>in</strong>cipal moments.<br />

Adeficiency of Euler’s equations is that the solutions yield the time variation of ω as seen from the bodyfixed<br />

reference frame axes, and not <strong>in</strong> the observers fixed <strong>in</strong>ertial coord<strong>in</strong>ate frame. Similarly the components<br />

of the external torques <strong>in</strong> the Euler equations are given with respect to the body-fixed axis system which<br />

implies that the orientation of the body is already known. Thus for non-zero external torques the problem<br />

cannot be solved until the the orientation is known <strong>in</strong> order to determ<strong>in</strong>e the components <br />

. However,<br />

these difficulties disappear when the external torques are zero, or if the motion of the body is known and it<br />

is required to compute the applied torques necessary to produce such motion.<br />

13.18 Lagrange equations of motion for rigid-body rotation<br />

The Euler equations of motion were derived us<strong>in</strong>g Newtonian concepts of torque and angular momentum.<br />

It is of <strong>in</strong>terest to derive the equations of motion us<strong>in</strong>g Lagrangian mechanics. It is convenient to use a<br />

generalized torque and assume that =0<strong>in</strong> the Lagrange-Euler equations. Note that the generalized<br />

force is a torque s<strong>in</strong>ce the correspond<strong>in</strong>g generalized coord<strong>in</strong>ate is an angle, and the conjugate momentum<br />

is angular momentum. If the body-fixed frame of reference is chosen to be the pr<strong>in</strong>cipal axes system, then,<br />

s<strong>in</strong>ce the <strong>in</strong>ertia tensor is diagonal <strong>in</strong> the pr<strong>in</strong>cipal axis frame, the k<strong>in</strong>etic energy is given <strong>in</strong> terms of the<br />

pr<strong>in</strong>cipal moments of <strong>in</strong>ertia as<br />

= 1 X<br />

2 (13.104)<br />

2<br />

<br />

Us<strong>in</strong>g the Euler angles as generalized coord<strong>in</strong>ates, then the Lagrange equation for the specific caseofthe<br />

coord<strong>in</strong>ate and <strong>in</strong>clud<strong>in</strong>g a generalized force gives<br />

which can be expressed as<br />

<br />

<br />

3X<br />

<br />

<br />

<br />

<br />

−<br />

˙ = (13.105)<br />

<br />

3X<br />

˙ − <br />

= (13.106)<br />

Equation 13104 gives<br />

<br />

= (13.107)<br />

<br />

Differentiat<strong>in</strong>g the angular velocity components <strong>in</strong> the body-fixed frame, equations (1386 − 1388) give<br />

1<br />

= ˙ s<strong>in</strong> cos − ˙ <br />

s<strong>in</strong> = 1<br />

2<br />

˙ = 2 =0<br />

˙<br />

2<br />

= − ˙ s<strong>in</strong> s<strong>in</strong> − ˙ <br />

cos = − 1 1<br />

˙ = 2 =0<br />

˙<br />

3<br />

=0 3<br />

=1<br />

˙<br />

Substitut<strong>in</strong>g these <strong>in</strong>to the Lagrange equation (13106) gives<br />

<br />

<br />

3 3 − 1 1 2 + 2 2 (− 1 )= 3 (13.108)


332 CHAPTER 13. RIGID-BODY ROTATION<br />

s<strong>in</strong>ce the and be 3 axes are col<strong>in</strong>ear. This can be rewritten as<br />

3 ˙ 3 − ( 1 − 2 ) 1 2 = 3 (13.109)<br />

Any axis could have been designated the be 3 axis, thus the above equation can be generalized to all three<br />

axes to give<br />

1 ˙ 1 − ( 2 − 3 ) 2 3 = 1 (13.110)<br />

2 ˙ 2 − ( 3 − 1 ) 3 1 = 2<br />

3 ˙ 3 − ( 1 − 2 ) 1 2 = 3<br />

These are the Euler’s equations given previously <strong>in</strong> (13103). Note that although ˙ 3 is the equation<br />

of motion for the coord<strong>in</strong>ate, this is not true for the φ and θ rotations which are not along the body-fixed<br />

1 and 2 axes as given <strong>in</strong> table 131.<br />

13.8 Example: Rotation of a dumbbell<br />

Consider the motion of the symmetric dumbbell shown <strong>in</strong> the adjacent figure. Let | 1 | = | 2 | = Let the<br />

body-fixed coord<strong>in</strong>ate system have its orig<strong>in</strong> at and symmetry axis be 3 be along the weightless shaft toward<br />

1 and v = ˆ 1 The angular momentum is given by<br />

L = X <br />

r × v <br />

Because L is perpendicular to the shaft, and L rotates around ω as the shaft rotates, let be 2 be along L<br />

L = 2 be 2<br />

If is the angle between ω and the shaft, the components of ω<br />

are<br />

1 = 0<br />

2 = s<strong>in</strong> <br />

3 = cos <br />

Assume that the pr<strong>in</strong>cipal moments of the dumbbell are<br />

1 = ( 1 + 2 ) 2<br />

2 = ( 1 + 2 ) 2<br />

3 = 0<br />

Thus the angular momentum is given by<br />

L<br />

O<br />

1 = 1 1 =0<br />

2 = 2 2 =( 1 + 2 ) 2 s<strong>in</strong> <br />

3 = 3 3 =0<br />

Rotation of a dumbbell.<br />

which is consistent with the angular momentum be<strong>in</strong>g along the be 2 axis.<br />

Us<strong>in</strong>g Euler’s equations, and assum<strong>in</strong>g that the angular velocity is constant, i.e. ˙ =0 then the components<br />

of the torque required to satisfy this motion are<br />

1 = − ( 1 + 2 ) 2 2 s<strong>in</strong> cos <br />

2 = 0<br />

3 = 0<br />

That is, this motion can only occur <strong>in</strong> the presence of the above applied torque which is <strong>in</strong> the direction<br />

− be 1 that is, mutually perpendicular to be 2 and be 3 . This torque can be written as N = ω × L.


13.19. HAMILTONIAN EQUATIONS OF MOTION FOR RIGID-BODY ROTATION 333<br />

13.19 Hamiltonian equations of motion for rigid-body rotation<br />

The Hamiltonian equations of motion are expressed <strong>in</strong> terms of the Euler angles plus their correspond<strong>in</strong>g<br />

canonical angular momenta ( ) <strong>in</strong> contrast to Lagrangian mechanics which is based on the<br />

Euler angles plus their correspond<strong>in</strong>g angular velocities ( ˙ ˙ ˙). The Hamiltonian approach is conveniently<br />

expressed <strong>in</strong> terms of a set of Andoyer-Deprit action-angle coord<strong>in</strong>ates that <strong>in</strong>clude the three Euler<br />

angles, specify<strong>in</strong>g the orientation of the body-fixed frame, plus the correspond<strong>in</strong>g three angles specify<strong>in</strong>g the<br />

orientation of the sp<strong>in</strong> frame of reference. This phase space approach[Dep67] can be employed for calculations<br />

of rotational motion <strong>in</strong> celestial mechanics that can <strong>in</strong>clude sp<strong>in</strong>-orbit coupl<strong>in</strong>g. This Hamiltonian<br />

approach is beyond the scope of the present textbook.<br />

13.20 Torque-free rotation of an <strong>in</strong>ertially-symmetric rigid rotor<br />

13.20.1 Euler’s equations of motion:<br />

There are many situations where one has rigid-body motion free<br />

of external torques, that is, N =0. The tumbl<strong>in</strong>g motion of a<br />

jugglers baton, a diver, a rotat<strong>in</strong>g galaxy, or a frisbee, are examples<br />

of rigid-body rotation. For torque-free rotation, the body<br />

will rotate about the center of mass, and thus the <strong>in</strong>ertia tensor<br />

with respect to the center of mass is required. An <strong>in</strong>ertiallysymmetric<br />

rigid body has two identical pr<strong>in</strong>cipal moments of<br />

<strong>in</strong>ertia with 1 = 2 6= 3 , and provides a simple example that<br />

illustrates the underly<strong>in</strong>g motion. The force-free Euler equations<br />

for the symmetric body <strong>in</strong> the body-fixed pr<strong>in</strong>cipal axis system<br />

are given by<br />

( 2 − 3 ) 2 3 − 1 ˙ 1 = 0 (13.111)<br />

( 3 − 1 ) 3 1 − 2 ˙ 2 = 0 (13.112)<br />

3 ˙ 3 = 0 (13.113)<br />

where 1 = 2 and =0apply.<br />

Note that for torque-free motion of an <strong>in</strong>ertially symmetric<br />

body equation 13113 implies that ˙ 3 =0 i.e. 3 is a constant<br />

of motion and thus is a cyclic variable for the symmetric rigid<br />

body.<br />

Equations 13111 and 13112 can be written as two coupled<br />

equations<br />

˙ 1 + Ω 2 = 0 (13.114)<br />

˙ 2 − Ω 1 = 0 (13.115)<br />

Figure 13.4: The force-free symmetric top<br />

angular velocity precesses on a conical<br />

trajectory about the body-fixed symmetry<br />

axis ˆ3.<br />

where the precession angular velocity Ω = ˙ with respect to the body-fixed frame is def<strong>in</strong>ed to be<br />

µ <br />

(3 − 1 )<br />

Ω ≡ ω 3<br />

1<br />

(13.116)<br />

Comb<strong>in</strong><strong>in</strong>g the time derivatives of equations 13114 and 13115 leads to two uncoupled equations<br />

¨ 1 + Ω 2 1 = 0 (13.117)<br />

¨ 2 + Ω 2 2 = 0 (13.118)<br />

These are the differential equations for a harmonic oscillator with solutions<br />

1 = cos Ω (13.119)<br />

2 = s<strong>in</strong> Ω


334 CHAPTER 13. RIGID-BODY ROTATION<br />

These equations describe a vector rotat<strong>in</strong>g <strong>in</strong> a circle of radius about an axis perpendicular to ˆ 3 that<br />

is, rotat<strong>in</strong>g <strong>in</strong> the ˆ 1 − ˆ 2 plane with angular frequency Ω = − ˙. Notethat<br />

2 1 + 2 2 = 2 (13.120)<br />

which is a constant. In addition 3 is constant, therefore the magnitude of the total angular velocity<br />

q<br />

|ω| = 2 1 + 2 2 + 2 3 = constant (13.121)<br />

The motion of the torque-free symmetric body is that the angular velocity ω precesses around the<br />

symmetry axis ˆ 3 of the body at an angle with a constant precession frequency Ω with respect to the<br />

body-fixed frame as shown <strong>in</strong> figure 134. Thus, to an observer on the body, ω traces out a cone around the<br />

body-fixed symmetry axis. Note from (13116) that the vectors Ωˆ 3 and 3ˆ 3 are parallel when Ω is positive,<br />

that is, 3 (oblate shape) and antiparallel if 3 (prolate shape).<br />

For the system considered, the orientation of the angular momentum vector L must be stationary <strong>in</strong> the<br />

space-fixed <strong>in</strong>ertial frame s<strong>in</strong>ce the system is torque free, that is, L is a constant of motion. Also we have<br />

that the projection of the angular momentum on the body-fixed symmetry axis is a constant of motion, that<br />

is, it is a cyclic variable. Thus<br />

3 = 3 3 = 1 3<br />

( 3 − 1 ) Ω (13.122)<br />

Understand<strong>in</strong>g the relation between the angular momentum and angular velocity is facilitated by consider<strong>in</strong>g<br />

another constant of motion for the torque-free symmetric rotor, namely the rotational k<strong>in</strong>etic energy.<br />

= 1 ω · L = constant (13.123)<br />

2<br />

S<strong>in</strong>ce L is a constant for torque-free motion, and also the magnitude of ω was shown to be constant, therefore<br />

the angle between these two vectors must be a constant to ensure that also rot = 1 2ω · L = constant. That<br />

is, ω precesses around L at a constant angle ( − ) such that the projection of ω onto L is constant. Note<br />

that<br />

ω × be 3 = 2 be 1 − 1 be 2 (13.124)<br />

and, for a symmetric rotor,<br />

L · ω × be 3 = 1 1 2 − 2 1 2 =0 (13.125)<br />

s<strong>in</strong>ce 1 = 2 for the symmetric rotor. Because L · ω × be 3 =0for a symmetric top then L ω and be 3 are<br />

coplanar.<br />

Figure 135 shows the geometry of the motion for both oblate and prolate axially-deformed bodies. To<br />

an observer <strong>in</strong> the space-fixed <strong>in</strong>ertial frame, the angular velocity ω traces out a cone that precesses with<br />

angular velocity Ω around the space fixed L axis called the space cone. For convenience, figure 135 assumes<br />

that L and the space-fixed <strong>in</strong>ertial frame ẑ axis are col<strong>in</strong>ear. The angular velocity ω also traces out the<br />

body cone as it precesses about the body-fixed ê 3 axis. S<strong>in</strong>ce L ω and be 3 are coplanar, then the ω vector is<br />

at the <strong>in</strong>tersection of the space and body cones as the body cone rolls around the space cone. That is, the<br />

space and body cones have one generatrix <strong>in</strong> common which co<strong>in</strong>cides with ω. Asshown<strong>in</strong>figure 135, for<br />

a needle the body cone appears to roll without slipp<strong>in</strong>g on the outside of the space cone at the precessional<br />

velocity of Ω = − By contrast, as shown <strong>in</strong> figure 135 for an oblate (disc-shaped) symmetric top the<br />

space cone rolls <strong>in</strong>side the body cone and the precession Ω is faster than .<br />

S<strong>in</strong>ce no external torques are act<strong>in</strong>g for torque-free motion, then the magnitude and direction of the total<br />

angular momentum are conserved. The description of the motion is simplified if L is taken to be along the<br />

space-fixed ẑ axis, then the Euler angle is the angle between the body-fixed basis vector ê 3 and space-fixed<br />

basis vector ẑ. Ifatsome<strong>in</strong>stant<strong>in</strong>thebodyframe,itisassumedthat be 2 is aligned <strong>in</strong> the plane of L ω<br />

and be 3 then<br />

1 =0 2 = s<strong>in</strong> 3 = cos (13.126)<br />

If is the angle between the angular velocity ω and the body-fixed ê 3 axis, then at the same <strong>in</strong>stant<br />

1 =0 2 = s<strong>in</strong> 3 = cos (13.127)


13.20. TORQUE-FREE ROTATION OF AN INERTIALLY-SYMMETRIC RIGID ROTOR 335<br />

Space cone<br />

3<br />

z<br />

L<br />

3<br />

z<br />

L<br />

Body cone<br />

Space cone<br />

Body cone<br />

2<br />

2<br />

1<br />

1<br />

(a)<br />

(b)<br />

Figure 13.5: Torque-free rotation of symmetric tops; (a) circular flat disk, (b) circular rod. The space-fixed<br />

and body-fixed cones are shown by f<strong>in</strong>e l<strong>in</strong>es. The space-fixed axis system is designated by the unit vectors<br />

(ˆx ŷ ẑ) and the body-fixed pr<strong>in</strong>cipal axis system by unit vectors (ˆ1 ˆ2 ˆ3)<br />

The components of the angular momentum also can be derived from L = I · ω to give<br />

1 = 1 1 =0 2 = 2 2 = 1 s<strong>in</strong> 3 = 3 3 = 3 cos (13.128)<br />

Equations 13126 and 13128 give two relations for the ratio 2<br />

3<br />

,thatis,<br />

2<br />

=tan = 1<br />

tan (13.129)<br />

3 3<br />

For a prolate spheroid 1 3 therefore while Ω and 3 have opposite signs.<br />

For a oblate spheroid 1 3 therefore while Ω and 3 havethesamesign.<br />

The sense of precession can be understood if the body cone rolls without slipp<strong>in</strong>g on the outside of the<br />

space cone with Ω <strong>in</strong> the opposite orientation to for the prolate case, while for the oblate case the space<br />

cone rolls <strong>in</strong>side the body cone with Ω and oriented <strong>in</strong> similar directions. Note from (13129) that =0<br />

if =0,thatisL ω and the 3 axis are aligned correspond<strong>in</strong>g to a pr<strong>in</strong>cipal axis. Similarly, =90 ◦ if<br />

=90 ◦ , then aga<strong>in</strong> L and ω are aligned correspond<strong>in</strong>g to them be<strong>in</strong>g pr<strong>in</strong>cipal axes.<br />

Lagrangian mechanics has been used to calculate the motion with respect to the body-fixed pr<strong>in</strong>cipal<br />

axis system. However, the motion needs to be known relative to the space-fixed <strong>in</strong>ertial frame where the<br />

motion is observed. This transformation can be done us<strong>in</strong>g the follow<strong>in</strong>g relation<br />

µ µ <br />

ê3<br />

ê3<br />

=<br />

+ ω × ê 3 = ω × ê 3 (13.130)<br />

<br />

<br />

<br />

<br />

s<strong>in</strong>ce the unit vector ê 3 is stationary <strong>in</strong> the body-fixed frame. The vector product of ω × ê 3 and ê 3 gives<br />

µ ê3<br />

ê 3 ×<br />

= ê 3 × ω × ê 3 =(ê 3 · ê 3 ) ω − (ê 3 · ω) ê 3 = ω − 3 ê 3<br />

<br />

<br />

therefore<br />

µ ê3<br />

ω = ê 3 ×<br />

+ 3 ê 3 (13.131)<br />

<br />

<br />

The angular momentum equals L = {I} ·ω. S<strong>in</strong>ce ê 3 × ¡ ê 3<br />

¢<br />

is perpendicular to the ê 3 axis, then<br />

for the case with 1 = 2 ,<br />

µ ê3<br />

L = 1 ê 3 ×<br />

+ 3 3 ê 3 (13.132)


336 CHAPTER 13. RIGID-BODY ROTATION<br />

Thus the angular momentum for a torque-free symmetric rigid rotor comprises two components, one be<strong>in</strong>g<br />

the perpendicular component that precesses around ê 3 , and the other is 3 .<br />

In the space-fixed frame assume that the ẑ axis is col<strong>in</strong>ear with L Then tak<strong>in</strong>g the scalar product of ê 3<br />

and L, us<strong>in</strong>g equation 13126 gives<br />

µ ê3<br />

3 = ê 3 · L = 1 ê 3 · ê 3 ×<br />

+ 3 3 ê 3 · ê 3 (13.133)<br />

<br />

<br />

The first term on the right is zero and thus equation 13133 and 13126 give<br />

3 = 3 3 = cos (13.134)<br />

Thetimedependenceoftherotationofthebody-fixed symmetry axis with respect to the space-fixed<br />

axis system can be obta<strong>in</strong>ed by tak<strong>in</strong>g the vector product ê 3 × L us<strong>in</strong>g equation 13132 and us<strong>in</strong>g equation<br />

24 to expand the triple vector product,<br />

à µ !<br />

ê3<br />

ê 3 × L = 1 ê 3 × ê 3 ×<br />

+ 3 3 ê 3 × ê 3 (13.135)<br />

= 1<br />

"Ãê 3 ·<br />

µ ê3<br />

<br />

<br />

<br />

<br />

<br />

!<br />

ê 3 − (ê 3 · ê 3 )<br />

µ #<br />

ê3<br />

+0<br />

<br />

<br />

s<strong>in</strong>ce (ê 3 × ê 3 )=0.Moreover(ê 3 · ê 3 )=1,andê 3 · ¡ ê 3<br />

¢<br />

<br />

=0 s<strong>in</strong>ce they are perpendicular, then<br />

<br />

µ ê3<br />

= L × ê 3 (13.136)<br />

<br />

<br />

1<br />

This equation shows that the body-fixed symmetry axis ê 3 precesses around the L where L is a constant<br />

of motion for torque-free rotation. The true rotational angular velocity ω <strong>in</strong> the space-fixed frame, given by<br />

equations 13131 can be evaluated us<strong>in</strong>g equation 13136 Remember<strong>in</strong>g that it was assumed that L is <strong>in</strong><br />

the ẑ direction, that is, L =ẑ then<br />

ω = ê 3 ×<br />

µ ê3<br />

+ 3 ê 3<br />

<br />

<br />

= µ cos <br />

ê 3 × (ẑ × ê 3 )+<br />

1<br />

= 1<br />

ẑ + cos <br />

3<br />

<br />

ê 3<br />

µ <br />

1 − 3<br />

ê 3 (13.137)<br />

1 3<br />

That is, the symmetry axis of the axially-symmetric rigid rotor makes an angle to the angular momentum<br />

vector ẑ and precesses around ẑ with a constant angular velocity 1<br />

while the axial sp<strong>in</strong> of the rigid body<br />

has a constant value 3<br />

. Thus, <strong>in</strong> the precess<strong>in</strong>g frame, the rigid body appears to rotate about its fixed<br />

symmetry axis with a constant angular velocity cos <br />

3<br />

− cos <br />

1<br />

= cos <br />

³<br />

1 − 3<br />

1 3<br />

´. The precession of the<br />

symmetry axis looks like a wobble superimposed on the sp<strong>in</strong>n<strong>in</strong>g motion about the body-fixed symmetry<br />

axis. The angular precession rate <strong>in</strong> the space-fixed frame can be deduced by us<strong>in</strong>g the fact that<br />

˙ s<strong>in</strong> = s<strong>in</strong> (13.138)<br />

Then us<strong>in</strong>g equation 13129 allows equation 13138 to be written as<br />

v<br />

" Ã<br />

u µ3 ! #<br />

2<br />

˙ = t<br />

1+ − 1 cos<br />

2 <br />

1<br />

(13.139)<br />

which gives the precession rate about the space-fixedaxis<strong>in</strong>termsoftheangularvelocity. Note that the<br />

precession rate ˙ if 3<br />

1<br />

1, that is, for oblate shapes, and ˙ if 3<br />

1<br />

1, that is, for prolate shapes.


13.20. TORQUE-FREE ROTATION OF AN INERTIALLY-SYMMETRIC RIGID ROTOR 337<br />

13.20.2 Lagrange equations of motion:<br />

It is <strong>in</strong>terest<strong>in</strong>g to compare the equations of motion for torque-free rotation of an <strong>in</strong>ertially-symmetric<br />

rigid rotor derived us<strong>in</strong>g Lagrange mechanics with that derived previously us<strong>in</strong>g Euler’s equations based on<br />

Newtonian mechanics. Assume that the pr<strong>in</strong>cipal moments about the fixed po<strong>in</strong>t of the symmetric top are<br />

1 = 2 6= 3 and that the k<strong>in</strong>etic energy equals the rotational k<strong>in</strong>etic energy, that is, it is assumed that the<br />

translational k<strong>in</strong>etic energy =0 Then the k<strong>in</strong>etic energy is given by<br />

Equations (1386 − 88) for the body-fixed frame give<br />

= 1 X<br />

2 = 1 2<br />

2 ¡<br />

1 <br />

2<br />

1 + 2 ¢ 1<br />

2 +<br />

2 3 2 3 (13.140)<br />

<br />

2 1 =<br />

³<br />

˙ s<strong>in</strong> s<strong>in</strong> + ˙ cos ´2<br />

= ˙2 s<strong>in</strong> 2 s<strong>in</strong> 2 +2˙ s<strong>in</strong> s<strong>in</strong> cos + ˙ 2 cos 2 (13.141)<br />

2 2 =<br />

³<br />

˙ s<strong>in</strong> cos − ˙ s<strong>in</strong> ´2<br />

= ˙2 s<strong>in</strong> 2 cos 2 − 2˙ s<strong>in</strong> s<strong>in</strong> cos + ˙ 2 s<strong>in</strong> 2 (13.142)<br />

Therefore<br />

and<br />

2 1 + 2 2 = ˙ 2 s<strong>in</strong> 2 + ˙ 2 (13.143)<br />

³<br />

2 ´2<br />

3 = ˙ cos + ˙<br />

(13.144)<br />

Therefore the k<strong>in</strong>etic energy is<br />

= 1 ³<br />

2 1 ˙2 s<strong>in</strong> 2 +<br />

2´<br />

˙ + 1 ³ ´2<br />

2 3 ˙ cos + ˙<br />

(13.145)<br />

S<strong>in</strong>ce the system is torque free, the scalar potential energy can be assumed to be zero, and then the<br />

Lagrangian equals<br />

= 1 ³<br />

2´<br />

2 1 ˙2 s<strong>in</strong> 2 + ˙ + 1 ³ ´2<br />

2 3 ˙ cos + ˙<br />

(13.146)<br />

The angular momentum about the space-fixed axis is conjugate to . From Lagrange’s equations<br />

˙ = =0 (13.147)<br />

<br />

that is, the angular momentum about the space-fixed axis, is a constant of motion given by<br />

= <br />

˙ = ¡ 1 s<strong>in</strong> 2 + 3 cos 2 ¢ ˙ + 3 ˙ cos = constant. (13.148)<br />

Similarly, the angular momentum about the body-fixed 3 axis is conjugate to From Lagrange’s equations<br />

˙ = =0 (13.149)<br />

<br />

that is, is a constant of motion given by<br />

= <br />

˙ = 3<br />

µ ˙ cos + ˙<br />

= 3 3 = constant (13.150)<br />

The above two relations derived from the Lagrangian can be solved to give the precession angular velocity<br />

˙ about the space-fixed ẑ axis<br />

˙ = − cos <br />

1 s<strong>in</strong> 2 (13.151)<br />

<br />

and the sp<strong>in</strong> about the body-fixed ˆ3 axis ˙ which is given by<br />

˙ = <br />

3<br />

− ( − cos )cos<br />

1 s<strong>in</strong> 2 <br />

(13.152)


338 CHAPTER 13. RIGID-BODY ROTATION<br />

S<strong>in</strong>ce and are constants of motion, then the precessional angular velocity ˙ about the space-fixed ẑ<br />

axis, and the sp<strong>in</strong> angular velocity ˙, which is the sp<strong>in</strong> frequency about the body-fixed ˆ3 axis, are constants<br />

that depend directly on 1 3 and <br />

There is one additional constant of motion available if no dissipative forces act on the system, that is,<br />

energy conservation which implies that the total energy<br />

= 1 ³<br />

2 1 ˙2 s<strong>in</strong> 2 +<br />

2´<br />

˙ + 1 ³ ´2<br />

2 3 ˙ cos + ˙<br />

(13.153)<br />

will be a constant of motion. But the second term on the right-hand side also is a constant of motion s<strong>in</strong>ce<br />

and 3 both are constants, that is<br />

1<br />

2 3 2 3 = 1 ³ ´2<br />

2 <br />

3 ˙ cos + ˙<br />

2 = = constant (13.154)<br />

3<br />

Thus energy conservation implies that the first term on the right-hand side also must be a constant given by<br />

1<br />

2 ¡<br />

1 <br />

2<br />

1 + 2 1<br />

³<br />

2´<br />

2¢<br />

=<br />

2 1 ˙2 s<strong>in</strong> 2 + ˙ = − 2 <br />

= constant (13.155)<br />

3<br />

These results are identical to those given <strong>in</strong> equations 13120 and 13121 which were derived us<strong>in</strong>g Euler’s<br />

equations. These results illustrate that the underly<strong>in</strong>g physics of the torque-free rigid rotor is more easily<br />

extracted us<strong>in</strong>g Lagrangian mechanics rather than us<strong>in</strong>g the Euler-angle approach of Newtonian mechanics.<br />

13.9 Example: Precession rate for torque-free rotat<strong>in</strong>g symmetric rigid rotor<br />

Table 132 lists the precession and sp<strong>in</strong> angular velocities, <strong>in</strong> the space-fixed frame, for torque-free rotation<br />

of three extreme symmetric-top geometries sp<strong>in</strong>n<strong>in</strong>g with constant angular momentum when the motion<br />

is slightly perturbed such that is at a small angle to the symmetry axis. Note that this assumes the<br />

perpendicular axis theorem, equation 1345 which states that for a th<strong>in</strong> lam<strong>in</strong>ae 1 + 2 = 3 giv<strong>in</strong>g, for a<br />

th<strong>in</strong> circular disk, 1 = 2 and thus 3 =2 1 <br />

Table 132:<br />

Precession and sp<strong>in</strong> rates for torque-free axial rotation of symmetric rigid rotors<br />

Rigid-body symmetric shape Pr<strong>in</strong>cipal moment ratio 3<br />

1<br />

Precession rate ˙ Sp<strong>in</strong> rate ˙<br />

Symmetric needle 0 0 <br />

Sphere 1 0<br />

Th<strong>in</strong> circular disk 2 2 −<br />

The precession angular velocity <strong>in</strong> the space frame ranges between 0 to 2 depend<strong>in</strong>g on whether the<br />

body-fixed sp<strong>in</strong> angular velocity is aligned or anti-aligned with the rotational frequency . For an extreme<br />

prolate spheroid 3<br />

1<br />

=0 the body-fixed sp<strong>in</strong> angular velocity Ω = − 3 which cancels the angular velocity<br />

of the rotat<strong>in</strong>g frame result<strong>in</strong>g <strong>in</strong> a zero precession angular velocity of the body-fixed ê 3 axis around the<br />

space-fixed frame. The sp<strong>in</strong> Ω =0<strong>in</strong> the body-fixed frame for the rigid sphere 3<br />

1<br />

=1 and thus the precession<br />

rate of the body-fixed ˆ 3 axis of the sphere around the space-fixed frame equals . For oblate spheroids and<br />

th<strong>in</strong> disks, such as a frisbee, 3<br />

1<br />

=2mak<strong>in</strong>g the body-fixed precession angular velocity Ω =+ which adds<br />

to the angular velocity and <strong>in</strong>creases the precession rate up to 2 as seen <strong>in</strong> the space-fixed frame. This<br />

illustrates that the sp<strong>in</strong> angular velocity can add constructively or destructively with the angular velocity 2<br />

2 In his autobiography Surely You’re Jok<strong>in</strong>g Mr Feynman, he wrote " I was <strong>in</strong> the [Cornell] cafeteria and some guy, fool<strong>in</strong>g<br />

around, throws a plate <strong>in</strong> the air. As the plate went up <strong>in</strong> the air I saw it wobble, and noticed that the red medallion of<br />

Cornell on the plate go<strong>in</strong>g around. It was pretty obvious to me that the medallion went around faster than the wobbl<strong>in</strong>g. I<br />

started to figure out the motion of the rotat<strong>in</strong>g plate. I discovered that when the angle is very slight, the medallion rotates<br />

twice as fast as the wobble rate. It came out of a very complicated equation! ". The quoted ratio (2 : 1) is <strong>in</strong>correct, it should<br />

be (1 : 2). Benjam<strong>in</strong> Chao <strong>in</strong> Physics Today of February 1989 speculated that Feynman’s error <strong>in</strong> <strong>in</strong>vert<strong>in</strong>g the factor of<br />

two might be "<strong>in</strong> keep<strong>in</strong>g with the spirit of the author and the book, another practical joke meant for those who do physics<br />

without experiment<strong>in</strong>g". He po<strong>in</strong>ted out that this story occurred on page 157 of a book of length 314 pages (1:2). Observe the<br />

dependence of the ratio of wobble to rotation angular velocities on the tilt angle .


13.21. TORQUE-FREE ROTATION OF AN ASYMMETRIC RIGID ROTOR 339<br />

13.21 Torque-free rotation of an asymmetric rigid rotor<br />

The Euler equations of motion for the case of torque-free rotation of an asymmetric (triaxial) rigid rotor<br />

about the center of mass, with pr<strong>in</strong>cipal moments of <strong>in</strong>ertia 1 6= 2 6= 3 lead to more complicated motion<br />

than for the symmetric rigid rotor. 3 The general features of the motion of the asymmetric rotor can be<br />

deduced us<strong>in</strong>g the conservation of angular momentum and rotational k<strong>in</strong>etic energy.<br />

Assum<strong>in</strong>g that the external torques are zero then the Euler<br />

equations of motion can be written as<br />

1 ˙ 1 = ( 2 − 3 ) 2 3 (13.156)<br />

2 ˙ 2 = ( 3 − 1 ) 3 1<br />

3 ˙ 3 = ( 1 − 2 ) 1 2<br />

S<strong>in</strong>ce = for =1 2 3, thenequation13156 gives<br />

2 3 ˙1 = ( 2 − 3 ) 2 3 (13.157)<br />

1 3 ˙2 = ( 3 − 1 ) 3 1<br />

1 2 ˙3 = ( 1 − 2 ) 1 2<br />

Multiply the first equation by 1 1 , the second by 2 2 and the<br />

third by 3 3 and sum, which gives<br />

1 2 3<br />

³<br />

1 ˙1 + 2 ˙2 + 3 ˙3´<br />

=0 (13.158)<br />

The bracket is equivalent to (2 1 + 2 2 + 2 3 )=0which implies<br />

that the total rotational angular momentum is a constant of<br />

motion as expected for this torque-free system, even though the<br />

<strong>in</strong>dividual components 1 2 3 may vary. That is<br />

2 1 + 2 2 + 2 3 = 2 (13.159)<br />

Figure 13.6: Rotation of an asymmetric<br />

rigid rotor. The dark l<strong>in</strong>es correspond to<br />

contours of constant total rotational k<strong>in</strong>etic<br />

energy T, which has an ellipsoidal<br />

shape, projected onto the angular momentum<br />

L sphere <strong>in</strong> the body-fixed frame.<br />

Note that equation 13159 is the equation of a sphere of radius .<br />

Multiply the first equation of 13157 by 1 , the second by 2 ,andthethirdby 3 ,andsumgives<br />

2 3 1 ˙1 + 1 3 2 ˙2 + 1 2 3 ˙3 =0 (13.160)<br />

Divide 13160 by 1 2 3 gives ( 2 1<br />

2 1<br />

+ 2 2<br />

2 2<br />

+ 2 3<br />

2 3<br />

)=0. This implies that the total rotational k<strong>in</strong>etic energy<br />

,givenby<br />

2 1<br />

+ 2 2<br />

+ 2 3<br />

= (13.161)<br />

2 1 2 2 2 3<br />

is a constant of motion as expected when there are no external torques and zero energy dissipation. Note<br />

that 13161 is the equation of an ellipsoid.<br />

Equations 13159 and 13161 both must be satisfied by the rotational motion for any value of the total<br />

angular momentum L and k<strong>in</strong>etic energy .Fig136 shows a graphical representation of the <strong>in</strong>tersection of<br />

the sphere and ellipsoid as seen <strong>in</strong> the body-fixed frame. The angular momentum vector L must follow<br />

the constant-energy contours given by where the -ellipsoids <strong>in</strong>tersect the -sphere, shown for the case where<br />

3 2 1 . Note that the precession of the angular momentum vector L follows a trajectory that has<br />

closed paths that circle around the pr<strong>in</strong>cipal axis with the smallest , thatis,ê 1 or the pr<strong>in</strong>cipal axis with<br />

the maximum , thatis,ê 3 . However, the angular momentum vector does not have a stable m<strong>in</strong>imum for<br />

precession around the <strong>in</strong>termediate pr<strong>in</strong>cipal moment of <strong>in</strong>ertia axis ê 2 . In addition to the precession, the<br />

angular momentum vector L executes nutation, that is a nodd<strong>in</strong>g of the angle <br />

For any fixed value of , the k<strong>in</strong>etic energy has upper and lower bounds given by<br />

2<br />

2 3<br />

≤ ≤ 2<br />

2 1<br />

(13.162)<br />

3 Similar discussions of the freely-rotat<strong>in</strong>g asymmetric top are given by Landau and Lifshitz [La60] and by Gregory [Gr06].


340 CHAPTER 13. RIGID-BODY ROTATION<br />

Thus, for a given value of when = m<strong>in</strong> = 2<br />

2 3<br />

the orientation of L <strong>in</strong> the body-fixed frame is either<br />

(0 0 +) or (0 0 −), that is, aligned with the ê 3 axis along which the pr<strong>in</strong>cipal moment of <strong>in</strong>ertia is largest.<br />

For slightly higher k<strong>in</strong>etic energy the trajectory of follows closed paths precess<strong>in</strong>g around ê 3 . When the<br />

k<strong>in</strong>etic energy = 2 2<br />

2 2<br />

the angular momentum vector follows either of the two th<strong>in</strong>-l<strong>in</strong>e trajectories each<br />

of which are a separatrix. These do not have closed orbits around ê 2 and they separate the closed solutions<br />

around either ê 3 or ê 1 For higher k<strong>in</strong>etic energy the precess<strong>in</strong>g angular momentum vector follows closed<br />

trajectories around ê 1 and becomes fully aligned with ê 1 at the upper-bound k<strong>in</strong>etic energy.<br />

Note that for the special case when 3 2 = 1 then the asymmetric rigid rotor equals the symmetric<br />

rigid rotor for which the solutions of Euler’s equations were solved exactly <strong>in</strong> chapter 1319. For the symmetric<br />

rigid rotor the -ellipsoid becomes a spheroid aligned with the symmetry axis and thus the <strong>in</strong>tersections<br />

with the -sphere lead to circular paths around the ê 3 body-fixed pr<strong>in</strong>cipal axis, while the separatrix circles<br />

the equator correspond<strong>in</strong>g to the ê 3 axis separat<strong>in</strong>g clockwise and anticlockwise precession about L 3 .This<br />

discussion shows that energy, plus angular momentum conservation, provide the general features of the<br />

solution for the torque-free symmetric top that are <strong>in</strong> agreement with those derived us<strong>in</strong>g Euler’s equations<br />

of motion<br />

13.22 Stability of torque-free rotation of an asymmetric body<br />

It is of <strong>in</strong>terest to extend the prior discussion to address the stability of an asymmetric rigid rotor undergo<strong>in</strong>g<br />

force-free rotation close to a pr<strong>in</strong>cipal axes, that is, when subject to small perturbations. Consider the case<br />

of a general asymmetric rigid body with 3 2 1 Let the system start with rotation about the ê 1 axis,<br />

that is, the pr<strong>in</strong>cipal axis associated with the moment of <strong>in</strong>ertia 1 Then<br />

ω = 1 be 1 (13.163)<br />

Consider that a small perturbation is applied caus<strong>in</strong>g the angular velocity vector to be<br />

where are very small. The Euler equations (13156) become<br />

ω = 1 be 1 + be 2 + be 3 (13.164)<br />

( 2 − 3 ) − 1 ˙ 1 = 0<br />

( 3 − 1 ) 1 − 2 ˙ = 0<br />

( 1 − 2 ) 1 − 3 ˙ = 0<br />

Assum<strong>in</strong>g that the product <strong>in</strong> the first equation is negligible, then ˙ 1 =0 that is, 1 is constant.<br />

The other two equations can be solved to give<br />

µ <br />

(3 − 1 )<br />

˙ =<br />

1 (13.165)<br />

2<br />

Take the time derivative of the first equation<br />

and substitute for ˙ gives<br />

The solution of this equation is<br />

where<br />

˙ =<br />

¨ =<br />

µ (1 − 2 )<br />

3<br />

1<br />

<br />

(13.166)<br />

µ (3 − 1 )<br />

2<br />

1<br />

<br />

˙ (13.167)<br />

µ <br />

(1 − 3 )( 1 − 2 )<br />

¨ +<br />

2 1 =0 (13.168)<br />

2 3<br />

() = Ω 1 + −Ω 1<br />

(13.169)<br />

Ω 1 = 1<br />

s<br />

( 1 − 3 )( 1 − 2 )<br />

2 3<br />

(13.170)


13.22. STABILITY OF TORQUE-FREE ROTATION OF AN ASYMMETRIC BODY 341<br />

Note that s<strong>in</strong>ce it was assumed that 3 2 1 then Ω 1 is real. The solution for () therefore represents a<br />

stable oscillatory motion with precession frequency Ω 1 The identical result is obta<strong>in</strong>ed for Ω 1 = Ω 1 = Ω 1 <br />

Thus the motion corresponds to a stable m<strong>in</strong>imum about the ê 1 axis with oscillations about the = =0<br />

m<strong>in</strong>imum with period.<br />

s<br />

( 1 − 3 )( 1 − 2 )<br />

Ω 1 = 1 (13.171)<br />

2 3<br />

Permut<strong>in</strong>g the <strong>in</strong>dices gives that for perturbations applied to rotation about either the 2 or 3 axes give<br />

precession frequencies<br />

s<br />

( 2 − 1 )( 2 − 3 )<br />

Ω 2 = 2 (13.172)<br />

1 3<br />

s<br />

( 3 − 2 )( 3 − 1 )<br />

Ω 3 = 3 (13.173)<br />

1 2<br />

S<strong>in</strong>ce 3 2 1 then Ω 1 and Ω 3 are real while Ω 2 is imag<strong>in</strong>ary. Thus, whereas rotation about either<br />

the 3 or the 1 axes are stable, the imag<strong>in</strong>ary solution about ê 2 corresponds to a perturbation <strong>in</strong>creas<strong>in</strong>g<br />

with time. Thus, only rotation about the largest or smallest moments of <strong>in</strong>ertia are stable. Moreover for<br />

the symmetric rigid rotor, with 1 = 2 6= 3 stability exists only about the symmetry axis ê 3 <strong>in</strong>dependent<br />

on whether the body is prolate or oblate. This result was implied from the discussion of energy and angular<br />

momentum conservation <strong>in</strong> chapter 1320. Friction was not <strong>in</strong>cluded <strong>in</strong> the above discussion. In the presence<br />

of dissipative forces, such as friction or drag, only rotation about the pr<strong>in</strong>cipal axis correspond<strong>in</strong>g to the<br />

maximum moment of <strong>in</strong>ertia is stable.<br />

Stability of rigid-body rotation has broad applications to rotation of satellites, molecules and nuclei.<br />

The first U.S. satellite, Explorer 1, was launched <strong>in</strong> 1958 with the rotation axis aligned with the cyl<strong>in</strong>drical<br />

axis which was the m<strong>in</strong>imum pr<strong>in</strong>cipal moment of <strong>in</strong>ertia. After a few hours the satellite started tumbl<strong>in</strong>g<br />

with <strong>in</strong>creas<strong>in</strong>g amplitude due to a flexible antenna dissipat<strong>in</strong>g and transferr<strong>in</strong>g energy to the perpendicular<br />

axis which had the largest moment of <strong>in</strong>ertia. Torque-free motion of a deformed rigid body is a ubiquitous<br />

phenomena <strong>in</strong> many branches of science, eng<strong>in</strong>eer<strong>in</strong>g, and sports as illustrated by the follow<strong>in</strong>g examples.<br />

13.10 Example: Tennis racquet dynamics<br />

A tennis racquet is an asymmetric body that exhibits the above rotational<br />

behavior. Assume that the head of a tennis racquet is a uniform<br />

th<strong>in</strong> circular disk of radius and mass whichisattachedtoacyl<strong>in</strong>drical<br />

handle of diameter = 10<br />

, length 2, and mass as shown <strong>in</strong><br />

the figure. The pr<strong>in</strong>ciple moments of <strong>in</strong>ertia about the three axes through<br />

the center-of-mass can be calculated by addition of the moments for the<br />

circular disk and the cyl<strong>in</strong>drical handle and us<strong>in</strong>g both the parallel-axis<br />

and the perpendicular-axis theorems.<br />

Axis Head Handle Racquet<br />

1<br />

1<br />

4 2 + 2 = 5 4 2 4 31<br />

32 12 2<br />

1<br />

2<br />

4 2 +0 = 1 4 2 1<br />

51<br />

2002 200 2<br />

1<br />

3<br />

2 2 + 2 = 3 2 2 4 3 2 17 6 2<br />

Note that 11 : 22 : 33 =25833 : 02550 : 28333. Insert<strong>in</strong>g these<br />

pr<strong>in</strong>ciple moments of <strong>in</strong>ertia <strong>in</strong>to equations 13171 − 13173 gives the<br />

follow<strong>in</strong>g precession frequencies<br />

Ω 1 = i0 8976 1 Ω 2 = 0 9056 2 Ω 3 = 0 9892 3<br />

M<br />

M<br />

Pr<strong>in</strong>cipal rotation axes for the<br />

center of mass of a tennis racket.<br />

The 1 and 2 -axes are <strong>in</strong> the<br />

plane of the racket head and the<br />

3 axis is perpendicular to the<br />

plane of the racket head.<br />

2<br />

1<br />

The imag<strong>in</strong>ary precession frequency Ω 1 about the 1 axis implies unstable rotation lead<strong>in</strong>g to tumbl<strong>in</strong>g<br />

whereas the m<strong>in</strong>imum moment 22 and maximum moment 33 imply stable rotation about the 2 and 3 axes.<br />

This rotational behavior is easily demonstrated by throw<strong>in</strong>g a tennis racquet and is called the tennis racquet<br />

theorem. The center of percussion, example 214 is another important <strong>in</strong>ertial property of a tennis racquet.


342 CHAPTER 13. RIGID-BODY ROTATION<br />

13.11 Example: Rotation of asymmetrically-deformed nuclei<br />

Some nuclei and molecules have average shapes that have significant asymmetric deformation lead<strong>in</strong>g to<br />

<strong>in</strong>terest<strong>in</strong>g quantal analogs of the rotational properties of an asymmetrically-deformed rigid body. The major<br />

difference between a quantal and a classical rotor is that the energies, and angular momentum are quantized,<br />

rather than be<strong>in</strong>g cont<strong>in</strong>uously variable quantities. Otherwise, the quantal rotors exhibit general features<br />

similar to the classical analog. Studies [Cli86] of the rotational behavior of asymmetrically-deformed nuclei<br />

exploit three aspects of classical mechanics, namely classical Coulomb trajectories, rotational <strong>in</strong>variants, and<br />

the properties of ellipsoidal rigid-bodies.<br />

Ellipsoidal deformation can be specified by the dimensions along each of the three pr<strong>in</strong>ciple axes. Bohr<br />

and Mottelson parameterized the ellipsoidal deformation <strong>in</strong> terms of three parameters, 0 which is the radius<br />

of the equivalent sphere, which is a measure of the magnitude of the ellipsoidal deformation from the sphere,<br />

and which specifies the deviation of the shape from axial symmetry. The ellipsoidal <strong>in</strong>tr<strong>in</strong>sic shape can be<br />

expressed <strong>in</strong> terms of the deviation from the equivalent sphere by the equation<br />

+2<br />

X<br />

( ) =( ) − 0 = 0<br />

=−2<br />

∗ 2 2 ( )<br />

()<br />

where ( ) is a Laplace spherical harmonic def<strong>in</strong>ed as<br />

( ) =<br />

s<br />

(2 +1) ( − )!<br />

4 ( + )! (cos ) −<br />

and (cos ) is an associated Legendre function of cos . Spherical harmonics are the angular portion of a<br />

set of solutions to Laplace’s equation. Represented <strong>in</strong> a system of spherical coord<strong>in</strong>ates, Laplace’s spherical<br />

harmonics ( ) are a specific set of spherical harmonics that form an orthogonal system. Spherical<br />

harmonics are important <strong>in</strong> many theoretical and practical applications.<br />

In the pr<strong>in</strong>cipal axis frame of the body, there are three non-zero quadrupole deformation parameters<br />

which can be written <strong>in</strong> terms of the deformation parameters where 20 = cos , 21 = 2−1 =0 and<br />

22 = 2−2 = √ 1 2<br />

s<strong>in</strong> Us<strong>in</strong>g these <strong>in</strong> equations () give the three semi-axis dimensions <strong>in</strong> the pr<strong>in</strong>cipal<br />

axis frame, (primed frame),<br />

r<br />

5<br />

=<br />

4 0 cos( − 2<br />

3 ) ()<br />

q<br />

q<br />

Note that for =0, then 1 = 2 = − 1 5<br />

2 4 5<br />

0 while 3 =+<br />

4 0, that is the body has prolate<br />

deformation with the symmetry axis along the 3 axis. The same prolate shape is obta<strong>in</strong>ed for = 2 3<br />

and<br />

= 4 3 with the prolate symmetry axes along the 1 and 2 axes respectively. For = 3 then 1 = 3 =<br />

q<br />

q<br />

+ 1 5<br />

2 4 5<br />

0 while 2 = −<br />

4 0, that is the body has oblate deformation with the symmetry axis along<br />

the 2 axis. The same oblate shape is obta<strong>in</strong>ed for = and = 5 3<br />

with the oblate symmetry axes along<br />

the 3 and 1 axes respectively. For other values of the shape is ellipsoidal.<br />

For the asymmetric deformed rigid body, the rotational Hamiltonian can be expressed <strong>in</strong> the form[Dav58]<br />

=<br />

3X<br />

=1<br />

|| 2<br />

4 2 s<strong>in</strong> 2 ( 0 − 2<br />

3 )<br />

where the rotational angular momentum is R The pr<strong>in</strong>cipal moments of <strong>in</strong>ertia are related by the triaxiality<br />

parameter 0 which they assumed is identical to the shape parameter . For axial symmetry the moment of<br />

<strong>in</strong>ertia about the symmetry axis is taken to be zero for a quantal system s<strong>in</strong>ce rotation of the potential well<br />

about the symmetry axis corresponds to no change <strong>in</strong> the potential well, or correspond<strong>in</strong>g rotation of the bound<br />

nucleons. That is, the nucleus is not a rigid body, the nucleons only rotate to the extent that the ellipsoidal<br />

potential well is cranked around such that the nucleons must follow the rotation of the potential well. In<br />

addition, vibrational modes coexist about the average asymmetric deformation, plus octupole deformation<br />

often coexists with the above quadrupole deformed modes.


13.23. SYMMETRIC RIGID ROTOR SUBJECT TO TORQUE ABOUT A FIXED POINT 343<br />

13.23 Symmetric rigid rotor subject to torque about a fixed po<strong>in</strong>t<br />

The motion of a symmetric top rotat<strong>in</strong>g <strong>in</strong> a gravitational field, with<br />

one po<strong>in</strong>t at a fixed location, is encountered frequently <strong>in</strong> rotational<br />

motion. Examples are the gyroscope and a child’s sp<strong>in</strong>n<strong>in</strong>g top.<br />

Rotation of a rigid rotor subject to torque about a fixed po<strong>in</strong>t, is a<br />

case where it is necessary to take the <strong>in</strong>ertia tensor with respect to<br />

the fixed po<strong>in</strong>t <strong>in</strong> the body, and not at the center of mass.<br />

Consider the geometry, shown <strong>in</strong> figure 137, where the symmetric<br />

top of mass is sp<strong>in</strong>n<strong>in</strong>g about a fixed tip that is displaced by<br />

adistance from the center of mass. The tip of the top is assumed<br />

to be at the orig<strong>in</strong> of both the space-fixed frame ( ) and the<br />

body-fixed frame (1 2 3) Assume that the translational velocity<br />

is zero and let the pr<strong>in</strong>cipal moments about the fixed po<strong>in</strong>t of the<br />

symmetric top be 1 = 2 6= 3 <br />

The Lagrange equations of motion can be derived assum<strong>in</strong>g that<br />

the k<strong>in</strong>etic energy equals the rotational k<strong>in</strong>etic energy, that is, it is<br />

assumed that the translational k<strong>in</strong>etic energy =0 Then the<br />

k<strong>in</strong>etic energy of an <strong>in</strong>ertially-symmetric rigid rotor can be derived<br />

for the torque-free symmetric top as given <strong>in</strong> equation 13145 to be<br />

= 1 X<br />

2 = 1 2<br />

2 ¡<br />

1 <br />

2<br />

1 + 2 ¢ 1<br />

2 +<br />

2 3 2 3 (13.174)<br />

<br />

= 1 ³<br />

2´<br />

2 1 ˙2 s<strong>in</strong> 2 + ˙ + 1 ³<br />

2 3 ˙ cos + ˙´2<br />

(13.175)<br />

3<br />

x<br />

Mg<br />

h<br />

z<br />

2<br />

1<br />

L<strong>in</strong>e of nodes<br />

Figure 13.7: Symmetric top sp<strong>in</strong>n<strong>in</strong>g<br />

about one fixed po<strong>in</strong>t.<br />

S<strong>in</strong>ce the potential energy is = cos then the Lagrangian<br />

equals<br />

= 1 ³<br />

2´<br />

2 1 ˙2 s<strong>in</strong> 2 + ˙ + 1 ³<br />

2 3 ˙ cos + ˙´2<br />

− cos (13.176)<br />

The angular momentum about the space-fixed axis is conjugate to . From Lagrange’s equations<br />

˙ = =0 (13.177)<br />

<br />

that is, is a constant of motion given by the generalized momentum<br />

= <br />

˙ = ¡ 1 s<strong>in</strong> 2 + 3 cos 2 ¢ ˙ + 3 ˙ cos = = constant (13.178)<br />

where is the angular momentum projection along the space-fixed axis.<br />

Similarly, the angular momentum about the body-fixed 3 axis is conjugate to From Lagrange’s equations<br />

˙ = =0 (13.179)<br />

<br />

that is, is a constant of motion given by the generalized momentum<br />

= µ<br />

˙ = 3<br />

˙ cos + ˙<br />

= 3 = constant (13.180)<br />

where 3 is the angular momentum projection along the body-fixed 3 axis. The above two relations can be<br />

solved to give the precessional angular velocity ˙ about the space-fixed axis<br />

˙ = − cos <br />

1 s<strong>in</strong> 2 <br />

= − 3 cos <br />

1 s<strong>in</strong> 2 <br />

(13.181)<br />

and the sp<strong>in</strong> angular velocity ˙ about the body-fixed 3 axis<br />

˙ = <br />

− ( − cos )cos<br />

3 1 s<strong>in</strong> 2 = 3<br />

− ( − 3 cos )cos<br />

3 1 s<strong>in</strong> 2 (13.182)<br />

<br />

S<strong>in</strong>ce and are constants of motion, i.e. 3 3 then these rotational angular velocities depend on only<br />

1 3 and <br />

y


344 CHAPTER 13. RIGID-BODY ROTATION<br />

There is one further constant of motion available if no frictional<br />

forces act on the system, that is, energy conservation. This implies<br />

that the total energy<br />

= 1 ³<br />

2 1 ˙2 s<strong>in</strong> 2 +<br />

˙<br />

2´<br />

+ 1 2 3<br />

³<br />

˙ cos + ˙´2<br />

+ cos (13.183)<br />

will be a constant of motion. But the middle term on the right-hand<br />

side also is a constant of motion<br />

1<br />

³ ´2<br />

2 <br />

3 ˙ cos + ˙<br />

2 = = 2 3<br />

= constant (13.184)<br />

3 3<br />

Thus energy conservation can be rewritten by def<strong>in</strong><strong>in</strong>g an energy 0<br />

where<br />

0 ≡ − 2 <br />

= 1 ³<br />

2´<br />

3 2 1 ˙2 s<strong>in</strong> 2 + ˙ +cos = constant (13.185)<br />

0<br />

This can be written as<br />

0 = 1 2 ˙ 2 1 + ( − cos ) 2<br />

2 1 s<strong>in</strong> 2 + cos (13.186)<br />

<br />

which can be expressed as<br />

Figure 13.8: Effective potential diagram<br />

for a sp<strong>in</strong>n<strong>in</strong>g symmetric top<br />

as a function of theta.<br />

where () is an effective potential<br />

0 = 1 2 1 ˙ 2 + () (13.187)<br />

() ≡ ( − cos ) 2<br />

2 1 s<strong>in</strong> 2 <br />

+ cos = ( − 3 cos ) 2<br />

2 1 s<strong>in</strong> 2 + cos (13.188)<br />

<br />

The effective potential () is shown <strong>in</strong> figure 138. It is clear that the motion of a symmetric top with<br />

effective energy 0 is conf<strong>in</strong>ed to angles 1 2 <br />

Note that the above result also is obta<strong>in</strong>ed if the Routhian is used, rather than the Lagrangian, as<br />

mentioned<strong>in</strong>chapter87, anddef<strong>in</strong>ed by equation (865). That is, the Routhian can be written as<br />

( ˙ ) = ˙ + ˙ − = ( ) − ( ˙) <br />

= − 1 2 ˙ 2 1 + ( − cos ) 2<br />

2 1 s<strong>in</strong> 2 + 2 <br />

+ cos (13.189)<br />

2 3<br />

The Routhian ( ˙ ) acts like a Hamiltonian for the ( ) and ( ) variables which are<br />

constants of motion, and thus are ignorable variables. The Routhian acts as the negative Lagrangian for the<br />

rema<strong>in</strong><strong>in</strong>g variable with rotational k<strong>in</strong>etic energy 1 2 1 ˙ 2 and effective potential energy <br />

= ( − cos ) 2<br />

2 1 s<strong>in</strong> 2 <br />

+ 2 <br />

3<br />

+ cos = ()+ 2 <br />

3<br />

The equation of motion describ<strong>in</strong>g the system <strong>in</strong> the rotat<strong>in</strong>g frame is given by one Lagrange equation<br />

<br />

( <br />

˙<br />

) − <br />

<br />

The negative sign of the Routhian cancels out when used <strong>in</strong> the Lagrange equation. Thus, <strong>in</strong> the rotat<strong>in</strong>g<br />

frame of reference, the system is reduced to a s<strong>in</strong>gle degree of freedom, the nutation angle with effective<br />

energy 0 given by equations 13186 − 13188.<br />

=0


13.23. SYMMETRIC RIGID ROTOR SUBJECT TO TORQUE ABOUT A FIXED POINT 345<br />

(a) (b) (c)<br />

Figure 13.9: Nutational motion of the body-fixed symmetry axis projected onto the space-fixed unit sphere.<br />

The three case are (a) ˙ never vanishes, (b) ˙ =0at = 2 (c) ˙ changes sign between 1 and 2 <br />

The motion of the symmetric top is simplest at the m<strong>in</strong>imum value of the effective potential curve, where<br />

0 = m<strong>in</strong> at which the nutation is restricted to a s<strong>in</strong>gle value = 0 The motion is a steady precession<br />

at a fixed angle of <strong>in</strong>cl<strong>in</strong>ation, that is, the “sleep<strong>in</strong>g top”. Solv<strong>in</strong>g for ¡ ¢<br />

<br />

= 0<br />

=0gives that<br />

"<br />

− cos = s<br />

#<br />

s<strong>in</strong> 2 0<br />

1 ± 1 − 4 1 cos 0<br />

2cos 0 2 (13.190)<br />

<br />

If 0 2<br />

then to ensure that the solution is real requires a m<strong>in</strong>imum value of the angular momentum on the<br />

body-fixed axis of 2 ≥ 4 1 cos 0 .If 0 2<br />

then there is no m<strong>in</strong>imum angular momentum projection<br />

on the body-fixed axis. There are two possible solutions to the quadratic relation correspond<strong>in</strong>g to either a<br />

slow or fast precessional frequency. Usually the slow precession is observed.<br />

For the general case, where 1 0 m<strong>in</strong> the nutation angle between the space-fixed and body-fixed 3<br />

axes varies <strong>in</strong> the range 1 2 This axis exhibits a nodd<strong>in</strong>g variation which is called nutation. Figure<br />

139 shows the projection of the body-fixed symmetry axis on the unit sphere <strong>in</strong> the space-fixed frame. Note<br />

that the observed nutation behavior depends on the relative sizes of and cos For certa<strong>in</strong> values, the<br />

precession ˙ changes sign between the two limit<strong>in</strong>g values of produc<strong>in</strong>g a loop<strong>in</strong>g motion as shown <strong>in</strong> figure<br />

139. Another condition is where the precession is zero for 2 produc<strong>in</strong>g a cusp at 2 as illustrated <strong>in</strong> figure<br />

139. This behavior can be demonstrated us<strong>in</strong>g the gyroscope or the symmetric top.<br />

13.12 Example: The Sp<strong>in</strong>n<strong>in</strong>g “Jack”<br />

The game “Jacks” is played us<strong>in</strong>g metal Jacks, each of which comprises<br />

six equal masses at the opposite ends of orthogonal axes of length<br />

Consider one jack sp<strong>in</strong>n<strong>in</strong>g around the body-fixed 3−axis with the lower<br />

mass at a fixed po<strong>in</strong>t on the ground, and with a steady precession around<br />

the space-fixed vertical axis with angle as shown. Assume that the<br />

body-fixed axes align with the arms of the jack.<br />

The pr<strong>in</strong>cipal moments of <strong>in</strong>ertia about one mass is given by the parallel<br />

axis theorem to be 2 = 1 =4 2 +6 2 =10 2 and 3 =4 2 .<br />

In the rotat<strong>in</strong>g body-fixed frame the torque due to gravity has components<br />

⎛<br />

N = ⎝<br />

6 s<strong>in</strong> s<strong>in</strong> <br />

6 s<strong>in</strong> cos <br />

0<br />

⎞<br />

⎠<br />

and the components of the angular velocity are<br />

⎛<br />

˙ s<strong>in</strong> s<strong>in</strong> + ˙ cos <br />

⎞<br />

ω = ⎝ ˙ s<strong>in</strong> cos − ˙ s<strong>in</strong><br />

˙ cos + ˙<br />

⎠<br />

Us<strong>in</strong>g Euler’s equations ( 13103) for the above components of and<br />

<strong>in</strong> the body-fixed frame, gives<br />

z<br />

O<br />

3<br />

S<br />

Jack comprises six bodies of<br />

mass at each end of<br />

orthogonal arms of length


346 CHAPTER 13. RIGID-BODY ROTATION<br />

10 ˙ 1 − 6 2 3 = 6 s<strong>in</strong> s<strong>in</strong> <br />

<br />

(a)<br />

10 ˙ 2 − 6 1 3 = 6 s<strong>in</strong> cos <br />

<br />

(b)<br />

4 ˙ 3 = 0 (c)<br />

Equation () relates the sp<strong>in</strong> about the 3 axis, the precession, and the angle to the vertical that is<br />

3 = ˙ cos + ˙ = Ω cos + = constant<br />

where ˙ ≡ is the sp<strong>in</strong> and ˙ ≡ Ω is the precession angular velocity.<br />

If the sp<strong>in</strong> axis is nearly vertical, ≈ 0 and thus s<strong>in</strong> ≈ and cos ≈ 1. Multiplyequation() × s<strong>in</strong> +<br />

() × cos and us<strong>in</strong>g the equations of the components of gives<br />

µ<br />

5¨ + 2Ω − 3Ω 2 − 3 <br />

=0<br />

<br />

The bracket must be positive to have stable s<strong>in</strong>usoidal oscillations.<br />

required for the jack to sp<strong>in</strong> about a stable vertical axis is given by.<br />

That is, the sp<strong>in</strong> angular velocity <br />

3Ω 2 + 3<br />

2Ω<br />

This example illustrates the conditions required for stable rotation of any axially-symmetric top.<br />

13.13 Example: The Tippe Top<br />

The Tippe Top comprises a section of a sphere, to<br />

which a short cyl<strong>in</strong>drical rod is mounted on the planar<br />

section, as illustrated. When the Tippe Top is spun on<br />

a horizontal surface this top exhibits the perverse behavior<br />

of transition<strong>in</strong>g from rotation with the spherical head<br />

rest<strong>in</strong>g on the horizontal surface, to flipp<strong>in</strong>g over such<br />

that it rotates rest<strong>in</strong>g on its elongated cyl<strong>in</strong>drical rod.<br />

The orientation of angular momentum rema<strong>in</strong>s roughly<br />

vertical as expected from conservation of angular momentum.<br />

This implies that the rotation with respect to<br />

the body-fixed axes must <strong>in</strong>vert as the top <strong>in</strong>verts. The<br />

center of mass is raised when the top <strong>in</strong>verts; the additional<br />

potential energy is provided by a reduction <strong>in</strong> the<br />

rotational k<strong>in</strong>etic energy.<br />

The Tippe Top behavior was first discovered <strong>in</strong> the<br />

1890’s but adequate solutions of the equations of motion<br />

have only been developed s<strong>in</strong>ce the 1950’s. S<strong>in</strong>ce the top<br />

precesses around the vertical axis, the po<strong>in</strong>t of contact is<br />

not on the symmetry axis of the top. Slid<strong>in</strong>g friction between<br />

the surface of the sp<strong>in</strong>n<strong>in</strong>g top and the horizontal<br />

surface provides a torque that causes the precession of<br />

the top to <strong>in</strong>crease and eventually flip up onto the cyl<strong>in</strong>drical<br />

peg. The Tippe Top is typical of many phenomena<br />

<strong>in</strong> physics where the underly<strong>in</strong>g physics pr<strong>in</strong>ciple can be<br />

3 axis<br />

ThegeometryoftheTippeTopofradius<br />

sp<strong>in</strong>n<strong>in</strong>g on a horizontal surface with slipp<strong>in</strong>g<br />

friction act<strong>in</strong>g between the top and the<br />

horizontal plane. The center of mass is a distance<br />

from the center of the spherical section along<br />

the axis of symmetry of the top.<br />

recognized but a detailed and rigorous solution can be complicated.<br />

The system has five degrees of freedom, which specify the location on the horizontal plane, plus the<br />

three Euler angles ( ). The paper by Cohen[Coh77] expla<strong>in</strong>s the motion <strong>in</strong> terms of Euler angles us<strong>in</strong>g<br />

the laboratory to body-fixed transformation relation. It shows that friction plays a pivotal role <strong>in</strong> the motion<br />

contrary to some earlier claims. Ciocci and Langerock[Cio07] used the Routhian to reduce the number<br />

z<br />

r<br />

a<br />

CG


13.24. THE ROLLING WHEEL 347<br />

of degrees of freedom from 5 to 2, namely which is the tilt angle, and 0 which is the orientation of the<br />

tilt. This Routhian is a Lagrangian <strong>in</strong> two dimension that was used to derive the equations of motion<br />

via the Lagrange Euler equation<br />

<br />

( <br />

˙<br />

<br />

( <br />

˙ 0<br />

) − <br />

<br />

= <br />

) − <br />

0 = 0<br />

where the 0 are generalized torques about the 2 angles that take <strong>in</strong>to account the slid<strong>in</strong>g frictional<br />

forces. This sophisticated Routhian reduction approach provides an exhaustive and ref<strong>in</strong>ed solution for the<br />

Tippe Top and confirms that slid<strong>in</strong>g friction plays a key role <strong>in</strong> the unusual behavior of the Tippe Top.<br />

13.24 The roll<strong>in</strong>g wheel<br />

As discussed <strong>in</strong> chapter 57 the roll<strong>in</strong>g wheel is a non-holonomic system that is simple <strong>in</strong> pr<strong>in</strong>ciple, but<br />

<strong>in</strong> practice the solution can be complicated, as illustrated by the Tippe Top. Chapter 1323 discussed the<br />

motion of a symmetric top rotat<strong>in</strong>g about a fixed po<strong>in</strong>t on the symmetry axis when subject to a torque. The<br />

roll<strong>in</strong>g wheel <strong>in</strong>volves rotation of a symmetric rigid body that is subject to torques. However, the po<strong>in</strong>t of<br />

contact of the wheel with a static plane is on the periphery of the wheel, and friction at the po<strong>in</strong>t of contact<br />

is assumed to ensure zero slip. Note that friction is necessary to ensure that the rotat<strong>in</strong>g object rolls without<br />

slipp<strong>in</strong>g, but the frictional force does no work for pure roll<strong>in</strong>g of an undeformable rigid wheel.<br />

Thecoord<strong>in</strong>atesystememployedisshown<strong>in</strong>Figure1310. For simplicity it is better to use a mov<strong>in</strong>g<br />

coord<strong>in</strong>ate frame (1 2 3) that is fixed to the orientation of the wheel with the orig<strong>in</strong> at the center of mass<br />

of the wheel, but this mov<strong>in</strong>g reference frame does not <strong>in</strong>clude the angular velocity ˙ of the disk about the<br />

3 axis. That is, the mov<strong>in</strong>g (1 2 3) frame has angular velocities<br />

1 = ˙ (13.191)<br />

2 = ˙ s<strong>in</strong> <br />

3 = ˙ cos <br />

The frame fixed <strong>in</strong> the rotat<strong>in</strong>g wheel must <strong>in</strong>clude the additional angular velocity of the disk ˙ about the<br />

ê 3 axis, that is<br />

Ω 1 = 1 = ˙ (13.192)<br />

Ω 2 = 2 = ˙ s<strong>in</strong> <br />

Ω 3 = 3 + ˙ = ˙ cos + ˙<br />

where Ω designates the angular velocity of the rotat<strong>in</strong>g disk, while ω designates the rotation of the mov<strong>in</strong>g<br />

frame (1 2 3).<br />

The pr<strong>in</strong>ciple moments of <strong>in</strong>ertia of a th<strong>in</strong> circular disk are related by the perpendicular axis theorem<br />

(chapter 139)<br />

1 + 2 = 3<br />

S<strong>in</strong>ce 1 = 2 for a uniform disk, therefore 3 =2 1 .<br />

Equation 1216 can be used to relate the vector forces F <strong>in</strong> the space-fixed frame to the rate of change<br />

of momenta <strong>in</strong> the mov<strong>in</strong>g frame (1 2 3) <br />

F = ṗ = ṗ + ω × p (13.193)<br />

This leads to the follow<strong>in</strong>g relations for the three components <strong>in</strong> the mov<strong>in</strong>g frame<br />

1 = ˙ 1 + 2 3 − 3 2 (13.194)<br />

2 − s<strong>in</strong> = ˙ 2 + 3 1 − 1 3<br />

3 − cos = ˙ 3 + 1 2 − 2 1


348 CHAPTER 13. RIGID-BODY ROTATION<br />

Figure 13.10: Uniform disk roll<strong>in</strong>g on a horizontal plane as viewed <strong>in</strong> the (a) fixed frame, and (b) roll<strong>in</strong>g<br />

disk frame. The space-fixed axis system is (x y z), while the mov<strong>in</strong>g reference frame (1 2 3) is centered at<br />

the center of mass of the disk with the 1 2 axes <strong>in</strong> the plane of the disk. The disk is rotat<strong>in</strong>g with a uniform<br />

angular velocity ˙ about the 3 axis and roll<strong>in</strong>g <strong>in</strong> the direction that is at an angle relative to the axis.<br />

where 1 2 3 are the reactive forces act<strong>in</strong>g shown <strong>in</strong> figure 1310.<br />

Similarly, the torques N <strong>in</strong> the space-fixed frame can be related to the rate of change of angular momentum<br />

by<br />

N = ˙L = ˙L + ω × L (13.195)<br />

where = I Ω . This leads to the follow<strong>in</strong>g relations for the three torque equations <strong>in</strong> the mov<strong>in</strong>g frame<br />

The roll<strong>in</strong>g constra<strong>in</strong>ts are<br />

1 = − 3 = 1 ˙Ω 1 + 3 Ω 3 2 − 2 Ω 2 3 (13.196)<br />

2 = 0 = 1 ˙Ω2 + 1 Ω 1 3 − 3 Ω 3 1<br />

3 = 1 = 3 ˙Ω3 + 2 Ω 2 1 − 1 Ω 1 2<br />

1 + Ω 3 = 0 (13.197)<br />

2 = 0<br />

3 − Ω 1 = 0<br />

where = . Comb<strong>in</strong><strong>in</strong>g equations 13194 13196 13197 gives<br />

¡<br />

1 + 2¢ ˙Ω1 + ¡ 3 + 2¢ 2 Ω 3 − 2 3 Ω 2 = −cos (13.198)<br />

1 ˙Ω2 + 1 3 Ω 1 − 3 1 Ω 3 = 0<br />

¡<br />

3 + 2¢ ˙Ω3 + 2 1 Ω 2 − ¡ 1 + 2¢ 2 Ω 1 = 0<br />

These are the torque equations about the po<strong>in</strong>t of contact .<br />

Introduction of equations 13191 and 13192 <strong>in</strong>to equation 13198 expresses the equations of motion <strong>in</strong><br />

terms of the Euler angles to be<br />

¡<br />

1 + 2¢ ¡ ¨ + 3 + 2¢ ³<br />

˙ s<strong>in</strong> ˙ cos + ˙´<br />

− 1 ˙2 s<strong>in</strong> cos = −cos (13.199)<br />

³ ´<br />

1¨ s<strong>in</strong> +21 ˙˙ cos − 3 ˙ ˙ cos + ˙ = 0<br />

¡<br />

3 + 2¢ ³¨ cos − ˙˙ s<strong>in</strong> + ¨´<br />

− 2 ˙ ˙ s<strong>in</strong> = 0


13.24. THE ROLLING WHEEL 349<br />

Equations 13199 are non-l<strong>in</strong>ear, and a closed-form solution is possible only for limited cases such as when<br />

=90 ◦ .<br />

Note that the above equations of motion also can be derived us<strong>in</strong>g Lagrangian mechanics know<strong>in</strong>g that<br />

= 1 2 ¡ 1 2 + 2 2 + 3<br />

2 ¢ 1 +<br />

2 ¡<br />

1 Ω<br />

2<br />

1 + Ω 2 1<br />

2¢<br />

+<br />

2 3Ω 2 3 − cos <br />

The differential equations of constra<strong>in</strong>t can be derived from equations 13197 to be<br />

− cos = 0<br />

− s<strong>in</strong> = 0<br />

Use of generalized forces plus the Lagrange-Euler equations (645) can be used to derive the equations of<br />

motion and solve for the components of the constra<strong>in</strong>t force 1 2 and 3 .<br />

13.14 Example: Tipp<strong>in</strong>g stability of a roll<strong>in</strong>g wheel<br />

A circular wheel roll<strong>in</strong>g <strong>in</strong> a vertical plane at high angular velocity <strong>in</strong>itially rolls <strong>in</strong> a straight l<strong>in</strong>e and<br />

rema<strong>in</strong>s vertical. However, below a certa<strong>in</strong> angular velocity, gyroscopic forces become weaker and the wheel<br />

will tip sideways and veer rapidly from the <strong>in</strong>itial direction. It is <strong>in</strong>terest<strong>in</strong>g to estimate the m<strong>in</strong>imum angular<br />

velocity of the disk such that it does not start to tip over sideways.<br />

Note that equations 13199 are satisfied for = 2 =0and ˙ = Ω 3 = constant. Assume a small<br />

disturbance causes the tilt angle to be = 2<br />

+ where is small and that is non-zero but small, that is<br />

˙ = ˙ and ˙ are small. Keep<strong>in</strong>g only terms to first order <strong>in</strong> the third of equations 13199 and <strong>in</strong>tegrat<strong>in</strong>g<br />

gives<br />

˙ cos + ˙ = Ω 3<br />

(a)<br />

The first two of equations 13198 become<br />

¡<br />

1 + 2¢ ¨ + ¡ 3 + 2¢ ˙Ω3 − = 0 (b)<br />

1¨ − 3 Ω 3 ˙ = 0 (c)<br />

Integrat<strong>in</strong>g equation () gives<br />

Insert<strong>in</strong>g () <strong>in</strong>to () gives<br />

¡<br />

1 + 2¢ ¨ +<br />

∙ ¡3<br />

+ 2¢ 3 Ω 2 3<br />

1<br />

˙ = 3Ω 3<br />

1<br />

(d)<br />

¸<br />

− =0<br />

Equation () has a stable oscillatory solution when the square bracket <strong>in</strong> positive, that is,<br />

Ω 2 1 <br />

3 <br />

3 ( 3 + 2 (f)<br />

)<br />

which gives the m<strong>in</strong>imum angular velocity required for stable roll<strong>in</strong>g motion. For angular velocity less than the<br />

m<strong>in</strong>imum, the square bracket <strong>in</strong> equation () is negative lead<strong>in</strong>g to an exponentially decay<strong>in</strong>g and divergent<br />

solution. For a uniform disk the perpendicular axis theorem gives 3 =2 1 = 1 2 2 for which equation ()<br />

gives<br />

Therefore the critical l<strong>in</strong>ear velocity of the wheel is<br />

Ω 2 3 <br />

<br />

3<br />

(e)<br />

(g)<br />

r<br />

<br />

= Ω 3 <br />

3<br />

(h)<br />

The bicycle wheel provides a common example of the tipp<strong>in</strong>g of a roll<strong>in</strong>g wheel. For the typical 035<br />

radius of a bicycle wheel, this gives a critical velocity of 107 =24. 4<br />

4 The stability of the bicycle is sensitive to the castor and other aspects of the steer<strong>in</strong>g geometry of the front wheel, <strong>in</strong><br />

addition to the gyroscopic effects. Excellent articles on this subject have been written by D.E.H. Jones Physics Today 23(4)<br />

(1970) 34, and also by J. Lowell & H.D. McKell, American Journal of Physics 50 (1982) 1106.


350 CHAPTER 13. RIGID-BODY ROTATION<br />

13.25 Dynamic balanc<strong>in</strong>g of wheels<br />

For rotat<strong>in</strong>g mach<strong>in</strong>ery It is crucial that rotors be both statically and dynamically balanced. Static balance<br />

means that the center of mass is on the axis of rotation. Dynamic balance means that the axis of rotation is<br />

a pr<strong>in</strong>cipal axis.<br />

For example, consider the symmetric rotor that has its symmetry axis at an angle to the axis of rotation.<br />

In this case the system is statically balanced s<strong>in</strong>ce the center of gravity is on the axis of rotation. However,<br />

the rotation axis is at an angle to the symmetry axis. This implies that the axle has to provide a torque<br />

to ma<strong>in</strong>ta<strong>in</strong> rotation that is not along a pr<strong>in</strong>cipal axis. If you distort the front wheel of your car by hitt<strong>in</strong>g it<br />

sideways aga<strong>in</strong>st the sidewalk curb, or if the wheel is not dynamically balanced, then you will f<strong>in</strong>d that the<br />

steer<strong>in</strong>g wheel can vibrate wildly at certa<strong>in</strong> speeds due to the torques caused by dynamic imbalance shak<strong>in</strong>g<br />

the steer<strong>in</strong>g mechanism. This can be especially bad when the rotation frequency is close to a resonant<br />

frequency of the suspension system. Insist that your automobile wheels are dynamically balanced when you<br />

change tires, static balanc<strong>in</strong>g will not elim<strong>in</strong>ate the dynamic imbalance forces. Another example is that the<br />

ailerons, rudder, and elevator on aircraft usually are dynamically balanced to stop the build up of oscillations<br />

that can couple to flex<strong>in</strong>g and flutter of the airframe which can lead to airframe failure.<br />

13.15 Example: Forces on the bear<strong>in</strong>gs of a rotat<strong>in</strong>g circular disk<br />

A homogeneous circular disk of mass , andradius,<br />

rotates with constant angular velocity about a body-fixed<br />

axis pass<strong>in</strong>g through the center of the circular disk as shown<br />

<strong>in</strong> the adjacent figure. The rotation axis is <strong>in</strong>cl<strong>in</strong>ed at an<br />

angle to the symmetry axis of the circular disk by bear<strong>in</strong>gs<br />

on both sides of the disk spaced a distance apart. Determ<strong>in</strong>e<br />

the forces on the bear<strong>in</strong>gs.<br />

Choose the body-fixed axes such that ˆ 3 is along the symmetry<br />

axis of the circular disk, and ˆ 1 po<strong>in</strong>ts <strong>in</strong> the plane of<br />

the disk symmetry axis and the rotation axis. These axes are<br />

the pr<strong>in</strong>cipal axes for which the <strong>in</strong>ertia tensor can be calculated<br />

to be<br />

I = 2<br />

4<br />

⎛<br />

⎝ 1 0 0 ⎞<br />

0 1 0 ⎠<br />

0 0 2<br />

Note that for this th<strong>in</strong> plane lam<strong>in</strong>ae disk 11 + 22 = 33 . Rotation of circular disk about an axis that<br />

The components of the angular velocity vector along the is at an angle to the symmetry axis of the<br />

three body-fixed axes are given by<br />

circular disk.<br />

ω =( s<strong>in</strong> 0cos )<br />

S<strong>in</strong>ce it is assumed that ˙ =0then substitut<strong>in</strong>g <strong>in</strong>to Euler’s equations (13103) gives the torques act<strong>in</strong>g to<br />

be<br />

1 = 3 =0<br />

2 = − 2 s<strong>in</strong> cos 1 4 2<br />

That is, the torque is <strong>in</strong> the ˆ 2 direction. Thus the forces on the bear<strong>in</strong>gs can be calculated s<strong>in</strong>ce N = r × F,<br />

thus<br />

| | = | 2|<br />

2<br />

= 2 2 s<strong>in</strong> 2<br />

<br />

16<br />

Estimate the size of these forces for the front wheel of your car travell<strong>in</strong>g at 70 m.p.h. if the rotation axis is<br />

displaced by 2 ◦ from the symmetry axis of the wheel.


13.26. ROTATION OF DEFORMABLE BODIES 351<br />

Figure 13.11: Forward two-and-a-half somersaults with two twists demonstrates unequivocally that a diver<br />

can <strong>in</strong>itiate cont<strong>in</strong>uous twist<strong>in</strong>g <strong>in</strong> midair. In the illustrated maneuver the diver does more than one full<br />

somersault before he starts to twist. To ma<strong>in</strong>ta<strong>in</strong> the twist<strong>in</strong>g the diver does not have to move his legs.[Fro80]<br />

13.26 Rotation of deformable bodies<br />

The discussion <strong>in</strong> this chapter has assumed that the rotat<strong>in</strong>g body is a rigid body. However, there is a<br />

broad and important class of problems <strong>in</strong> classical mechanics where the rotat<strong>in</strong>g body is deformable that<br />

leads to <strong>in</strong>trigu<strong>in</strong>g new phenomena. The classic example is the cat, which, if dropped upside down with zero<br />

angular momentum, is able to distort its body plus tail <strong>in</strong> order to rotate such that it lands on its feet <strong>in</strong><br />

spite of the fact that there are no external torques act<strong>in</strong>g and thus the angular momentum is conserved.<br />

Another example is the high diver do<strong>in</strong>g a forward two—and-a-half somersault with two twists.[Fro80] Once<br />

the diver leaves the board then the total angular momentum must be conserved s<strong>in</strong>ce there are no external<br />

torques act<strong>in</strong>g on the system. The diver beg<strong>in</strong>s a somersault by rotat<strong>in</strong>g about a horizontal axis which is a<br />

pr<strong>in</strong>cipal axis that is perpendicular to the axis of his body pass<strong>in</strong>g through his hips. Initially the angular<br />

momentum, and angular velocity, are parallel and po<strong>in</strong>t perpendicular to the symmetry axis. Initially the<br />

diver goes <strong>in</strong>to a tuck which greatly reduces his moment of <strong>in</strong>ertia along the axis of his somersault which<br />

concomitantly <strong>in</strong>creases his angular velocity about this axis and he performs one full somersault prior to<br />

<strong>in</strong>itiat<strong>in</strong>g twist<strong>in</strong>g. Then the diver twists its body and moves its arms to destroy the axial symmetry of his<br />

body which changes the direction of the pr<strong>in</strong>cipal axes of the <strong>in</strong>ertia tensor. This causes the angular velocity<br />

to change <strong>in</strong> both direction and magnitude such that the angular momentum rema<strong>in</strong>s conserved. The angular<br />

velocity now is no longer parallel to the angular momentum result<strong>in</strong>g <strong>in</strong> a component along the length of<br />

the body caus<strong>in</strong>g it to twist while somersault<strong>in</strong>g. This twist<strong>in</strong>g motion will cont<strong>in</strong>ue until the symmetry<br />

of the diver’s body is restored which is done just before enter<strong>in</strong>g the water. By skilled tim<strong>in</strong>g, and body<br />

movement, the diver restores the symmetry of his body to the optimum orientation for enter<strong>in</strong>g the water.<br />

Such phenomena <strong>in</strong>volv<strong>in</strong>g deformable bodies are important to motion of ballet dancers, jugglers, astronauts<br />

<strong>in</strong> space, and satellite motion. The above rotational phenomena would be impossible if the cat or diver were<br />

rigid bodies hav<strong>in</strong>g a fixed <strong>in</strong>ertia tensor. Calculation of the dynamics of the motion of deformable bodies<br />

is complicated and beyond the scope of this book, but the concept of a time dependent transformation of<br />

the <strong>in</strong>ertia tensor underlies the subsequent motion. The theory is complicated s<strong>in</strong>ce it is difficult even to<br />

quantify what corresponds to rotation as the body morphs from one shape to another. Further <strong>in</strong>formation<br />

on this topic can be found <strong>in</strong> the literature. [Fro80]


352 CHAPTER 13. RIGID-BODY ROTATION<br />

13.27 Summary<br />

This chapter has <strong>in</strong>troduced the important, topic of rigid-body rotation which has many applications <strong>in</strong><br />

physics, eng<strong>in</strong>eer<strong>in</strong>g, sports, etc.<br />

Inertia tensor The concept of the <strong>in</strong>ertia tensor was <strong>in</strong>troduced where the 9 components of the <strong>in</strong>ertia<br />

tensor are given by<br />

Z Ã Ã 3X<br />

! !<br />

= (r 0 ) 2 − <br />

(1314)<br />

<br />

Ste<strong>in</strong>er’s parallel-axis theorem<br />

11 ≡ 11 + ¡¡ 2 1 + 2 2 + 2 3¢<br />

11 − 2 1¢<br />

= 11 + ¡ 2 2 + 2 ¢<br />

3<br />

(1343)<br />

relates the <strong>in</strong>ertia tensor about the center-of-mass to that about parallel axis system not through the center<br />

of mass.<br />

Diagonalization of the <strong>in</strong>ertia tensor about any po<strong>in</strong>t was used to f<strong>in</strong>d the correspond<strong>in</strong>g Pr<strong>in</strong>cipal axes<br />

of the rigid body.<br />

Angular momentum The angular momentum L for rigid-body rotation is expressed <strong>in</strong> terms of the<br />

<strong>in</strong>ertia tensor and angular frequency by<br />

⎛<br />

L= ⎝ ⎞ ⎛<br />

11 12 13<br />

21 22 23<br />

⎠ · ⎝ ⎞<br />

1<br />

2<br />

⎠ = {I}·ω<br />

(1356)<br />

31 32 33 3<br />

Rotational k<strong>in</strong>etic energy The rotational k<strong>in</strong>etic energy is<br />

⎛<br />

= 1 ¡ ¢<br />

1 2 3 · ⎝ ⎞ ⎛<br />

11 12 13<br />

21 22 23<br />

⎠ · ⎝ ⎞<br />

1<br />

2<br />

⎠<br />

2<br />

31 32 33 3<br />

(1372)<br />

≡ T = 1 2 ω ·{I}·ω = 1 2 ω · L (1373)<br />

Euler angles The Euler angles relate the space-fixed and body-fixed pr<strong>in</strong>cipal axes. The angular velocity<br />

ω expressed <strong>in</strong> terms of the Euler angles has components for the angular velocity <strong>in</strong> the body-fixed axis system<br />

(1 2 3)<br />

1 = ˙ 1 + ˙ 1 + 1 = ˙ s<strong>in</strong> s<strong>in</strong> + ˙ cos (1386)<br />

2 = ˙ 2 + ˙ 2 + 2 = ˙ s<strong>in</strong> cos − ˙ s<strong>in</strong> (1387)<br />

3 = ˙ 3 + ˙ 3 + 3 = ˙ cos + ˙ (1388)<br />

Similarly, the components of the angular velocity for the space-fixed axis system ( ) are<br />

= ˙ cos + ˙ s<strong>in</strong> s<strong>in</strong> (1389)<br />

= ˙ s<strong>in</strong> − ˙ s<strong>in</strong> cos (1390)<br />

= ˙ + ˙ cos (1391)<br />

Rotational <strong>in</strong>variants The powerful concept of the rotational <strong>in</strong>variance of scalar properties was <strong>in</strong>troduced.<br />

Important examples of rotational <strong>in</strong>variants are the Hamiltonian, Lagrangian, and Routhian.<br />

Euler equations of motion for rigid-body motion The dynamics of rigid-body rotational motion was<br />

explored and the Euler equations of motion were derived us<strong>in</strong>g both Newtonian and Lagrangian mechanics.<br />

<br />

1 = 1<br />

<br />

1 − ( 2 − 3 ) 2 3 (13103)<br />

<br />

2 = 2<br />

<br />

2 − ( 3 − 1 ) 3 1<br />

<br />

3 = 3<br />

<br />

3 − ( 1 − 2 ) 1 2


13.27. SUMMARY 353<br />

Lagrange equations of motion for rigid-body motion The Euler equations of motion for rigid-body<br />

motion, given <strong>in</strong> equation 13103 were derived us<strong>in</strong>g the Lagrange-Euler equations.<br />

Torque-free motion of rigid bodies The Euler equations and Lagrangian mechanics were used to study<br />

torque-free rotation of both symmetric and asymmetric bodies <strong>in</strong>clud<strong>in</strong>g discussion of the stability of torquefree<br />

rotation.<br />

Rotat<strong>in</strong>g symmetric body subject to a torque The complicated motion exhibited by a symmetric top,<br />

that is sp<strong>in</strong>n<strong>in</strong>g about one fixed po<strong>in</strong>t and subject to a torque, was <strong>in</strong>troduced and solved us<strong>in</strong>g Lagrangian<br />

mechanics.<br />

The roll<strong>in</strong>g wheel The non-holonomic motion of roll<strong>in</strong>g wheels was <strong>in</strong>troduced, as well as the importance<br />

of static and dynamic balanc<strong>in</strong>g of rotat<strong>in</strong>g mach<strong>in</strong>ery..<br />

Rotation of deformable bodies<br />

bodies was <strong>in</strong>troduced.<br />

The complicated non-holonomic motion <strong>in</strong>volv<strong>in</strong>g rotation of deformable


354 CHAPTER 13. RIGID-BODY ROTATION<br />

Problems<br />

1. A hollow spherical shell has a mass and radius .<br />

(a) Calculate the <strong>in</strong>ertia tensor for a set of coord<strong>in</strong>ates whose orig<strong>in</strong> is at the center of mass of the shell.<br />

(b) Now suppose that the shell is roll<strong>in</strong>g without slipp<strong>in</strong>g toward a step of height , where.Theshell<br />

has a l<strong>in</strong>ear velocity . What is the angular momentum of the shell relative to the tip of the step?<br />

(c) The shell now strikes the tip of the step <strong>in</strong>elastically (so that the po<strong>in</strong>t of contact sticks to the step,<br />

but the shell can still rotate about the tip of the step). What is the angular momentum of the shell<br />

immediately after contact?<br />

(d) F<strong>in</strong>ally, f<strong>in</strong>d the m<strong>in</strong>imum velocity which enables the shell to surmount the step. Express your result <strong>in</strong><br />

terms of and .<br />

2. The vectors ˆ, ˆ, andˆ constitute a set of orthogonal right-handed axes. The vectors ˆ +ˆ − 2ˆ, −ˆ +ˆ, and<br />

ˆ +ˆ +ˆ are also perpendicular to one another.<br />

(a) Write out the set of direction cos<strong>in</strong>es relat<strong>in</strong>g the new axes to the old.<br />

(b) How are the Eulerian angles def<strong>in</strong>ed? Describe this transformation by a set of Eulerian angles.<br />

3. A torsional pendulum consists of a vertical wire attached to a mass which can rotate about the vertical axis.<br />

Consider three torsional pendula which consist of identical wires from which identical homogeneous solid cubes<br />

are hung. One cube is hung from a corner, one from midway along an edge, and one from the middle of a face<br />

as shown. What are the ratios of the periods of the three pendula?<br />

4. A dumbbell comprises two equal po<strong>in</strong>t masses connected by a massless rigid rod of length 2 which is<br />

constra<strong>in</strong>ed to rotate about an axle fixed to the center of the rod at an angle asshown<strong>in</strong>thefigure. The<br />

center of the rod is at the orig<strong>in</strong> of the coord<strong>in</strong>ates, the axle along the -axis, and the dumbbell lies <strong>in</strong> the<br />

− plane at =0. The angular velocity is a constant <strong>in</strong> time and is directed along the axis.<br />

a) Calculate all elements of the <strong>in</strong>ertia tensor. Be sure to specify the coord<strong>in</strong>ate system used.<br />

b) Us<strong>in</strong>g the calculated <strong>in</strong>ertia tensor f<strong>in</strong>d the angular momentum of the dumbbell <strong>in</strong> the laboratory frame as<br />

a function of time.<br />

c) Us<strong>in</strong>g the equation = × , calculate the angular momentum and show that it it is equal to the answer<br />

of part (b).<br />

d) Calculate the torque on the axle as a function of time.<br />

e) Calculate the k<strong>in</strong>etic energy of the dumbbell.<br />

x<br />

O<br />

z


13.27. SUMMARY 355<br />

5. A heavy symmetric top has a mass with the center of mass a distance from the fixed po<strong>in</strong>t about which<br />

it sp<strong>in</strong>s and 1 = 2 6= 3 . The top is precess<strong>in</strong>g at a steady angular velocity Ω about the vertical space-fixed<br />

axis. What is the m<strong>in</strong>imum sp<strong>in</strong> 0 about the body-fixed symmetry axis, that is, the 3 axis assum<strong>in</strong>g that<br />

the 3 axis is <strong>in</strong>cl<strong>in</strong>ed at an angle = with respect to the vertical axis. Solve the problem at the <strong>in</strong>stant<br />

when the 3 1 axes all are <strong>in</strong> the same plane as shown <strong>in</strong> the figure.<br />

z<br />

3<br />

O<br />

x<br />

1<br />

6. Consider an object with the center of mass is at the orig<strong>in</strong> and <strong>in</strong>ertia tensor,<br />

⎛<br />

⎞<br />

12 −12 0<br />

= ⎝ −12 12 0 ⎠<br />

0 0 1<br />

(a) Determ<strong>in</strong>e the pr<strong>in</strong>cipal moments of <strong>in</strong>ertia and the pr<strong>in</strong>cipal axes. Guess the object.<br />

(b) Determ<strong>in</strong>e the rotation matrix and compute † . Do the diagonal elements match with your results<br />

from (a)? Note: columns of are eigenvectors of .<br />

(c) Assume = √ 2<br />

(ˆ +ˆ). Determ<strong>in</strong>e <strong>in</strong> the rotat<strong>in</strong>g coord<strong>in</strong>ate system. Are and <strong>in</strong> the same<br />

direction? What does this mean?<br />

(d) Repeat (c) for = √ 2<br />

(ˆ − ˆ). Whatisdifferent and why?<br />

(e) For which case will there be a non-zero torque required?<br />

(f) Determ<strong>in</strong>e the rotational k<strong>in</strong>etic energy for the case = √<br />

2<br />

(ˆ − ˆ)?<br />

7. Consider a wheel (solid disk) of mass and radius . The wheel is subject to angular velocities = ˆ<br />

where ˆ is normal to the surface and = ˆ.<br />

(a) Choose a set of pr<strong>in</strong>cipal axes by observation.<br />

(b) Determ<strong>in</strong>e the angular velocities and angular momentum along the pr<strong>in</strong>cipal axes. Note: 1 = 1 2 2 and<br />

2 = 3 = 1 4 2 .<br />

(c) Determ<strong>in</strong>e the torque.<br />

(d) Determ<strong>in</strong>e the rotation matrix that rotates the fixed coord<strong>in</strong>ate system to the body coord<strong>in</strong>ate system.<br />

8. Determ<strong>in</strong>e the pr<strong>in</strong>cipal moments of <strong>in</strong>ertia of an ellipsoid given by the equation,<br />

2<br />

2 + 2<br />

2 + 2<br />

2 =1


356 CHAPTER 13. RIGID-BODY ROTATION<br />

9. Determ<strong>in</strong>e the pr<strong>in</strong>cipal moments of <strong>in</strong>ertia of a sphere of radius with a cavity of radius located from the<br />

center of the sphere.<br />

10. Three equal masses form³<br />

the vertices ´ of an ³ equilateral triangle ´ of side length . Assume that the masses<br />

are located at<br />

³0 0 √<br />

3´<br />

, 0 2 − <br />

2 √ ,and 0 − 3<br />

2 − <br />

2 √ , such that the center-of-mass is located at the<br />

3<br />

orig<strong>in</strong>.<br />

(a) Determ<strong>in</strong>e the pr<strong>in</strong>cipal moments of <strong>in</strong>ertia and pr<strong>in</strong>cipal axes.<br />

Now consider the same system rotated 45 ◦ about the ˆ-axis. Assume that the masses are located at<br />

³<br />

´ ³<br />

´<br />

− <br />

2 √ <br />

2 2 √ − <br />

2 2 √ <br />

,and<br />

3<br />

2 √ − <br />

2 2 √ − <br />

2 2 √ , respectively.<br />

3<br />

(b) Determ<strong>in</strong>e the pr<strong>in</strong>cipal moments of <strong>in</strong>ertia and pr<strong>in</strong>cipal axes.<br />

(c) Could you have answered (b) without explicitly determ<strong>in</strong><strong>in</strong>g the <strong>in</strong>ertia tensor? How?<br />

³0 0 √<br />

3´<br />

,<br />

11. Calculate the moments of <strong>in</strong>ertia 1 2 3 for a homogeneous cone of mass whose height is and whose<br />

base has a radius Choose the 3 -axis along the symmetry axis of the cone.<br />

a) Choose the orig<strong>in</strong> at the apex of the cone, and calculate the elements of the <strong>in</strong>ertia tensor.<br />

b) Make a transformation such that the center of mass of the cone is the orig<strong>in</strong> and f<strong>in</strong>d the pr<strong>in</strong>cipal moments<br />

of <strong>in</strong>ertia.<br />

12. Four masses, all of mass lie <strong>in</strong> the − plane at positions ( ) =( 0) (− 0) (0 +2) (0 −2)<br />

Thesearejo<strong>in</strong>edbymasslessrodstoformarigidbody<br />

(a) F<strong>in</strong>d the <strong>in</strong>ertial tensor, us<strong>in</strong>g the axes as a reference system. Exhibit the tensor as a matrix.<br />

(b) Consider a direction given by the unit vector ˆ that lies equally between the positive axes; that is<br />

it makes equal angles with these three directions. F<strong>in</strong>d the moment of <strong>in</strong>ertia for rotation about this ˆ axis.<br />

(c) Given that at a certa<strong>in</strong> time the angular velocity vector lies along the above direction ˆ, f<strong>in</strong>d, for that<br />

<strong>in</strong>stant, the angle between the angular momentum vector and ˆ<br />

13. A homogeneous cube, each edge of which has a length <strong>in</strong>itially is <strong>in</strong> a position of unstable equilibrium with<br />

one edge of the cube <strong>in</strong> contact with a horizontal plane. The cube then is given a small displacement caus<strong>in</strong>g<br />

it to tip over and fall. Show that the angular velocity of the cube when one face strikes the plane is given by<br />

2 = ³√<br />

2 − 1´<br />

<br />

where = 3 2 if the edge cannot slide on the plane, and where = 12 5<br />

if slid<strong>in</strong>g can occur without friction.<br />

14. A symmetric body moves without the <strong>in</strong>fluence of forces or torques. Let 3 be the symmetry axis of the body<br />

and be along 0 3 . The angle between and 3 is . Let and <strong>in</strong>itially be <strong>in</strong> the 2 − 3 plane. What is<br />

the angular velocity of the symmetry axis about <strong>in</strong> terms of 1 3 and ?<br />

15. Consider a th<strong>in</strong> rectangular plate with dimensions by and mass Determ<strong>in</strong>e the torque necessary to<br />

rotate the th<strong>in</strong> plate with angular velocity about a diagonal. Expla<strong>in</strong> the physical behavior for the case when<br />

= .


Chapter 14<br />

Coupled l<strong>in</strong>ear oscillators<br />

14.1 Introduction<br />

Chapter 3 discussed the behavior of a s<strong>in</strong>gle l<strong>in</strong>early-damped l<strong>in</strong>ear oscillator subject to a harmonic force.<br />

No account was taken for the <strong>in</strong>fluence of the s<strong>in</strong>gle oscillator on the driver for the case of forced oscillations.<br />

Many systems <strong>in</strong> nature comprise complicated free or forced oscillations of coupled-oscillator systems. Examples<br />

of coupled oscillators are; automobile suspension systems, electronic circuits, electromagnetic fields,<br />

musical <strong>in</strong>struments, atoms bound <strong>in</strong> a crystal, neural circuits <strong>in</strong> the bra<strong>in</strong>, networks of pacemaker cells <strong>in</strong><br />

the heart, etc. Energy can be transferred back and forth between coupled oscillators as the motion evolves.<br />

It is possible to describe the motion of coupled l<strong>in</strong>ear oscillators <strong>in</strong> terms of a sum over <strong>in</strong>dependent normal<br />

coord<strong>in</strong>ates, i.e. normal modes, even though the motion may be very complicated. These normal modes<br />

are constructed from the orig<strong>in</strong>al coord<strong>in</strong>ates <strong>in</strong> such a way that the normal modes are uncoupled. The<br />

topic of f<strong>in</strong>d<strong>in</strong>g the normal modes of coupled oscillator systems is a ubiquitous problem encountered <strong>in</strong> all<br />

branches of science and eng<strong>in</strong>eer<strong>in</strong>g. As discussed <strong>in</strong> chapter 3 oscillatory motion of non-l<strong>in</strong>ear systems<br />

can be complicated. Fortunately most oscillatory systems are approximately l<strong>in</strong>ear when the amplitude of<br />

oscillation is small. This discussion assumes that the oscillation amplitudes are sufficiently small to ensure<br />

l<strong>in</strong>earity.<br />

14.2 Two coupled l<strong>in</strong>ear oscillators<br />

Consider the two-coupled l<strong>in</strong>ear oscillator, shown <strong>in</strong> figure<br />

141, which comprises two identical masses each connected to<br />

fixed locations by identical spr<strong>in</strong>gs hav<strong>in</strong>g a force constant<br />

. A spr<strong>in</strong>g with force constant 0 couples the two oscillators.<br />

The equilibrium lengths of the outer two spr<strong>in</strong>gs are <br />

while that of the coupl<strong>in</strong>g spr<strong>in</strong>g is 0 . The problem is simplified<br />

by restrict<strong>in</strong>g the motion to be along the l<strong>in</strong>e connect<strong>in</strong>g<br />

the masses and assum<strong>in</strong>g fixed endpo<strong>in</strong>ts. The small displacements<br />

of 1 and 2 are taken to be 1 and 2 with respect to<br />

the equilibrium positions and + 0 respectively. The restor<strong>in</strong>g<br />

force on 1 is − 1 − 0 ( 1 − 2 ) while the restor<strong>in</strong>g force<br />

on 2 is − 2 − 0 ( 2 − 1 ) This coupled double-oscillator<br />

system exhibits basic features of coupled l<strong>in</strong>ear oscillator systems.<br />

Assum<strong>in</strong>g 1 = 2 = then the equations of motion<br />

are<br />

¨ 1 +( + 0 ) 1 − 0 2 = 0 (14.1)<br />

¨ 2 +( + 0 ) 2 − 0 1 = 0<br />

x 1<br />

x 2<br />

m<br />

cm<br />

Figure 14.1: Two coupled l<strong>in</strong>ear oscillators.<br />

The equilibrium spr<strong>in</strong>g-lengths are for the<br />

outer spr<strong>in</strong>gs and 0 for the coupl<strong>in</strong>g spr<strong>in</strong>g.<br />

The displacement from the stable locations<br />

are given by 1 and 2 . The separation between<br />

the two masses is and the location of<br />

the center-of-mass is .<br />

m<br />

Assume that the motion for these coupled equations is oscil-<br />

357


358 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

latory with a solution of the form<br />

1 = 1 (14.2)<br />

2 = 2 <br />

where the constants may be complex to take <strong>in</strong>to account both the magnitude and phase. Substitut<strong>in</strong>g<br />

these possible solutions <strong>in</strong>to the equations of motion gives<br />

− 2 1 +( + 0 ) 1 − 0 2 = 0 (14.3)<br />

− 2 2 +( + 0 ) 2 − 0 1 = 0<br />

Collect<strong>in</strong>g terms, and cancell<strong>in</strong>g the common exponential factor,<br />

gives<br />

¡<br />

+ 0 − 2¢ 1 − 0 2 = 0 (14.4)<br />

¡<br />

+ 0 − 2¢ 2 − 0 1 = 0<br />

The existence of a non-trivial solution of these two simultaneous<br />

equations requires that the determ<strong>in</strong>ant of the coefficients of<br />

1 and 2 must vanish, that is<br />

¯ + 0 − 2 − 0<br />

− 0 + 0 − <br />

¯¯¯¯ 2 =0 (14.5)<br />

The expansion of this secular determ<strong>in</strong>ant yields<br />

¡<br />

+ 0 − 2¢ 2<br />

− 02 =0 (14.6)<br />

Solv<strong>in</strong>g for gives r<br />

+ 0 ± <br />

=<br />

(14.7)<br />

<br />

That is, there are two characteristic frequencies (or eigenfrequencies)<br />

for the system<br />

1 =<br />

r<br />

+2<br />

0<br />

<br />

(14.8)<br />

r <br />

2 =<br />

(14.9)<br />

<br />

Figure 14.2: Displacement of each of two<br />

S<strong>in</strong>ce superposition applies for these l<strong>in</strong>ear equations, then the coupled l<strong>in</strong>ear harmonic oscillators with<br />

general solution can be written as a sum of the terms that account =4and 0 =1<strong>in</strong> relative units.<br />

for the two possible values of .<br />

Figure 142 shows the solutions for a case where =4and 0 =1 <strong>in</strong> arbitrary units, with the q <strong>in</strong>itial<br />

6<br />

condition that 2 = and 1 = ˙ 1 = ˙ 2 =0. The two characteristic frequencies are 1 =<br />

<br />

q<br />

and<br />

4<br />

2 =<br />

<br />

. The characteristic beats phenomenon is exhibited where the envelope over one complete cycle of<br />

the low frequency encompasses several higher frequency oscillations. That is, the solution is<br />

2 () = £<br />

<br />

1 + −1 + 2 + − 2 ¤ ∙µ ¸ ∙µ ¸<br />

1 + 2 1 − 2<br />

= cos cos<br />

(14.10)<br />

4<br />

2<br />

2<br />

while<br />

1 () = £<br />

<br />

1 + −1 − 2 − −2¤ ∙µ ¸ ∙µ ¸<br />

1 + 2 1 − 2<br />

= s<strong>in</strong> s<strong>in</strong><br />

(14.11)<br />

4<br />

2<br />

2<br />

The energy <strong>in</strong> the two-coupled oscillators flows back and forth between the coupled oscillators as illustrated<br />

<strong>in</strong> figure 142.<br />

A better understand<strong>in</strong>g of the energy flow occurr<strong>in</strong>g between the two coupled oscillators is given by<br />

us<strong>in</strong>g a ( 1 2 ) configuration-space plot, shown <strong>in</strong> figure 143 The flow of energy occurr<strong>in</strong>g between the two<br />

coupled oscillators can be represented by choos<strong>in</strong>g normal-mode coord<strong>in</strong>ates 1 and 2 that are rotated by<br />

45 ◦ with respect to the spatial coord<strong>in</strong>ates ( 1 2 ). These normal-mode coord<strong>in</strong>ates ( 1 2 ) correspond to<br />

the two normal modes of the coupled double-oscillator system.


14.3. NORMAL MODES 359<br />

14.3 Normal modes<br />

The normal modes of the two-coupled oscillator system are<br />

obta<strong>in</strong>ed by a transformation to a pair of normal coord<strong>in</strong>ates<br />

( 1 2 ) that are <strong>in</strong>dependent and correspond to the two normal<br />

modes. The pair of normal coord<strong>in</strong>ates for this case are<br />

that is<br />

1 ≡ 1 − 2 (14.12)<br />

2 ≡ 1 + 2<br />

1 = 1 2 ( 2 + 1 ) (14.13)<br />

2 = 1 2 ( 2 − 1 )<br />

Substitute these <strong>in</strong>to the equations of motion (141), gives<br />

¡ <br />

1 + ¢<br />

2 +( +2 0 ) 1 + 0 2 = 0 (14.14)<br />

¡ <br />

1 − ¢<br />

2 +( +2 0 ) 1 − 0 2 = 0<br />

Add<strong>in</strong>g and subtract<strong>in</strong>g these two equations gives<br />

¨ 1 +( +2 0 ) 1 = 0 (14.15)<br />

¨ 2 + 2 = 0<br />

Note that the two coord<strong>in</strong>ates 1 and 2 are uncoupled and therefore<br />

are <strong>in</strong>dependent. The solutions of these equations are<br />

Figure 14.3: Motion of two coupled harmonic<br />

oscillators <strong>in</strong> the ( 1 2 ) spatial<br />

configuration space and <strong>in</strong> terms of the<br />

normal modes ( 1 2 ). Initial conditions<br />

are 2 = 1 = ˙ 1 = ˙ 2 =0<br />

1 () = 1 + 1 + 1 − −1 (14.16)<br />

2 () = 2 + 2 + 2 − −2<br />

where 1 corresponds to angular frequencies 1 ,and 2 corresponds to 2 .Thetwocoord<strong>in</strong>ates 1 and 2 are<br />

called the normal coord<strong>in</strong>ates and the two solutions are the normal modes with correspond<strong>in</strong>g<br />

angular frequencies, 1 and 2 .<br />

The ( 1 2 ) axes of the two normal modes correspond to a<br />

rotation of 45 ◦ <strong>in</strong> configuration space, figure 143. The <strong>in</strong>itial<br />

conditions chosen correspond to 1 = − 2 and thus both modes<br />

are excited with equal <strong>in</strong>tensity. Note that there are 5 lobes along<br />

the 2 axis versus 4 lobes along the 1 axis reflect<strong>in</strong>g the ratio<br />

of the eigenfrequencies 1 and 2 Also note that the diamond<br />

shape of the motion <strong>in</strong> the ( 1 2 ) configuration space illustrates<br />

that the extrema amplitudes for 2 are a maximum when 1 is<br />

zero, and vise versa. This is equivalent to the statement that<br />

the energies <strong>in</strong> the two modes are coupled with the energy for<br />

the first oscillator be<strong>in</strong>g a maximum when the energy is a m<strong>in</strong>imum<br />

for the second oscillator, and vise versa. By contrast, <strong>in</strong><br />

the ( 1 2 ) configuration space, the motion is bounded by a rectangle<br />

parallel to the ( 1 2 ) axes reflect<strong>in</strong>g the fact that the<br />

extrema amplitudes, and correspond<strong>in</strong>g energies, for the 1 normal<br />

mode are constant and <strong>in</strong>dependent of the motion for the 2<br />

normal mode, and vise versa. The decoupl<strong>in</strong>g of the two normal<br />

modes is best illustrated by consider<strong>in</strong>g the case when only one<br />

of these two normal modes is excited. For the <strong>in</strong>itial conditions<br />

1 (0) = − 2 (0) and 1 (0) = − 2 (0) then 2 () =0 That is,<br />

only the 1 () normal mode is excited with frequency 1 which<br />

corresponds to motion conf<strong>in</strong>ed to the 1 axis of figure 143<br />

1<br />

Antisymmetric mode<br />

(out of phase)<br />

2<br />

Symmetric mode<br />

(<strong>in</strong> phase)<br />

Figure 14.4: Normal modes for two coupled<br />

oscillators.


360 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

As shown <strong>in</strong> figure 144, 1 () is the antisymmetric mode <strong>in</strong> which the two masses oscillate out of phase<br />

such as to keep the center of mass of the two masses stationary. For the <strong>in</strong>itial conditions 1 (0) = 2 (0) <br />

and 1 (0) = 2 (0) then 1 () =0 that is, only the 2 () normal mode is excited. The 2 () normal mode<br />

is the symmetric mode where the two masses oscillate <strong>in</strong> phase with frequency 2 ; it corresponds to motion<br />

along the 2 axis For the symmetric phase, both masses move together lead<strong>in</strong>g to a constant extension of<br />

the coupl<strong>in</strong>g spr<strong>in</strong>g. As a result the frequency 2 of the symmetric mode 2 () is lower than the frequency<br />

1 of the asymmetric mode 1 () That is, the asymmetric mode is stiffer s<strong>in</strong>ce all three spr<strong>in</strong>gs provide<br />

active restor<strong>in</strong>g forces, compared to the symmetric mode where the coupl<strong>in</strong>g spr<strong>in</strong>g is uncompressed. In<br />

general, for attractive forces the lowest frequency always occurs for the mode with the highest symmetry.<br />

14.4 Center of mass oscillations<br />

Transform<strong>in</strong>g the coord<strong>in</strong>ates <strong>in</strong>to the center of mass of the two oscillat<strong>in</strong>g masses elucidates an <strong>in</strong>terest<strong>in</strong>g<br />

feature of the normal modes for the two-coupled l<strong>in</strong>ear oscillator. As illustrated <strong>in</strong> figure 141, the centerof-mass<br />

coord<strong>in</strong>ate for the two mass system is<br />

while the relative separation distance is<br />

2 = + 1 + + 0 + 2 =2 + 0 + 2<br />

That is, the two normal modes are<br />

=( + 0 + 2 ) − ( + 1 )= 0 − 1<br />

1 = 0 − (14.17)<br />

2 = 2 − 2 − 0<br />

q<br />

+2 0<br />

<br />

The 1 mode, which has angular frequency 1 = corresponds to an oscillations of the relative<br />

separation , while the center-of-mass location is stationary. By contrast, the 2 mode, with angular<br />

frequency 2 = p <br />

corresponds to an oscillation of the center of mass with the relative separation <br />

be<strong>in</strong>g a constant.<br />

Figure 145 illustrates the decoupled center-of-mass<br />

, and relative motions for both normal modes of<br />

the coupled double-oscillator system. The difference <strong>in</strong><br />

angular frequencies and amplitudes is readily apparent.<br />

It is of <strong>in</strong>terest to consider the special case where the<br />

spr<strong>in</strong>g constant =0for the two outside spr<strong>in</strong>gs. Then<br />

q<br />

2 0<br />

<br />

the angular frequencies are 1 = and 2 =0for<br />

the two normal modes. When =0the 2 mode is a<br />

spurious center-of-mass mode s<strong>in</strong>ce it corresponds to an<br />

oscillation with 2 =0<strong>in</strong> spite of the fact that there<br />

are no forces act<strong>in</strong>g on the center of mass. That is, the<br />

center-of-mass momentum must be a constant of motion.<br />

This spurious center-of-mass oscillation is a consequence<br />

of measur<strong>in</strong>g the displacements ( 1 2 ) with respect to<br />

an arbitrary external reference that is not related to the<br />

center of mass of the coupled system. Spurious centerof-mass<br />

modes are encountered frequently <strong>in</strong> many-body<br />

coupled oscillator systems such as molecules and nuclei.<br />

In such cases it is necessary to project out the center-ofmass<br />

motion to elim<strong>in</strong>ate such spurious solutions as will<br />

be discussed later.<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

0 1 2 3 4 5 6 7 8 9 10<br />

Figure 14.5: Time dependence of the center-ofmass<br />

and relative separation for two coupled<br />

l<strong>in</strong>ear oscillators assum<strong>in</strong>g spr<strong>in</strong>g constants<br />

of =4 and 0 = .<br />

t<br />

R<br />

r<br />

cm


that is r <br />

= 0 (1 − ) (14.24)<br />

14.5. WEAK COUPLING 361<br />

14.5 Weak coupl<strong>in</strong>g<br />

If one of the two coupled l<strong>in</strong>ear oscillator masses is held fixed, then the other free mass will oscillate with a<br />

frequency.<br />

r<br />

+ <br />

0<br />

0 =<br />

(14.18)<br />

<br />

The effect of coupl<strong>in</strong>g of the two oscillators is to split the degeneracy of the frequency for each mass to<br />

r r r +2<br />

0<br />

+ <br />

1 =<br />

0<br />

<br />

0 =<br />

2 =<br />

(14.19)<br />

Thus the degeneracy is broken, and the two normal modes have frequencies straddl<strong>in</strong>g the s<strong>in</strong>gle-oscillator<br />

frequency.<br />

It is <strong>in</strong>terest<strong>in</strong>g to consider the case where the coupl<strong>in</strong>g is weak because this situation occurs frequently<br />

<strong>in</strong> nature. The coupl<strong>in</strong>g is weak if the coupl<strong>in</strong>g constant 0 Then<br />

r<br />

+2<br />

0<br />

1 =<br />

<br />

= r <br />

<br />

√<br />

1+4 (14.20)<br />

where<br />

Thus<br />

≡ 0<br />

1 (14.21)<br />

2<br />

r <br />

1 ≈ (1 + 2) (14.22)<br />

<br />

The natural frequency of a s<strong>in</strong>gle oscillator was shown to be<br />

r r + <br />

0 <br />

0 =<br />

≈ (1 + ) (14.23)<br />

<br />

Thus the frequencies for the normal modes for weak coupl<strong>in</strong>g<br />

can be written as<br />

0<br />

1<br />

r <br />

1 = (1 + 2)<br />

<br />

≈ 0 (1 − )(1+2) ≈ 0 (1 + ) (14.25)<br />

n=2<br />

2<br />

while<br />

2 =<br />

r <br />

≈ 0 (1 − ) (14.26)<br />

That is the two solutions are split equally spaced about the<br />

s<strong>in</strong>gle uncoupled oscillator value given by 0 = ≈<br />

p <br />

<br />

q<br />

+ 0<br />

<br />

(1 + ). Note that the s<strong>in</strong>gle uncoupled oscillator frequency<br />

0 depends on the coupl<strong>in</strong>g strength 0 .<br />

This splitt<strong>in</strong>g of the characteristic frequencies is a feature<br />

exhibited by many systems of identical oscillators where<br />

half of the frequencies are shifted upwards and half downward.<br />

If is odd, then the central frequency is unshifted as<br />

illustrated for the case of =3. An example of this behavior<br />

is the Zeeman effect where the magnetic field couples the<br />

atomic motion result<strong>in</strong>g <strong>in</strong> a hyperf<strong>in</strong>e splitt<strong>in</strong>g of the energy<br />

levels as illustrated.<br />

1<br />

0<br />

2<br />

n=3<br />

Figure 14.6: Normal-mode frequencies for<br />

n=2 and n=3 weakly-coupled oscillators.<br />

3


362 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

There are myriad examples <strong>in</strong>volv<strong>in</strong>g weakly-coupled oscillators <strong>in</strong> many aspects of the natural world.<br />

The example of collective modes <strong>in</strong> nuclear physics, illustrated <strong>in</strong> example 1413, is typical of applications to<br />

physics, while there are many examples applied to musical <strong>in</strong>struments, acoustics, and eng<strong>in</strong>eer<strong>in</strong>g. Weaklycoupled<br />

oscillators are a dom<strong>in</strong>ant theme throughout biology as illustrated by congregations of synchronously<br />

flash<strong>in</strong>g fireflies, crickets that chirp <strong>in</strong> unison, an audience clapp<strong>in</strong>g at the end of a performance, networks<br />

of pacemaker cells <strong>in</strong> the heart, <strong>in</strong>sul<strong>in</strong>-secret<strong>in</strong>g cells <strong>in</strong> the pancreas, and neural networks <strong>in</strong> the bra<strong>in</strong> and<br />

sp<strong>in</strong>al cord that control rhythmic behaviors such as breath<strong>in</strong>g, walk<strong>in</strong>g, and eat<strong>in</strong>g. Synchronous motion of<br />

a large number of weakly-coupled oscillators often leads to large collective motion of weakly-coupled systems<br />

as discussed <strong>in</strong> chapter 1412<br />

14.1 Example: The Grand Piano<br />

Damper<br />

Str<strong>in</strong>g<br />

Bridge<br />

Hitchp<strong>in</strong><br />

Hammer<br />

P<strong>in</strong> block<br />

Jack<br />

Key<br />

Soundboard<br />

Ribs<br />

Schematic diagram of the action for a grand piano, <strong>in</strong>clud<strong>in</strong>g the str<strong>in</strong>gs, bridge and sound<strong>in</strong>g board. Note<br />

that there are either two or three parallel str<strong>in</strong>gs per note that are hit by a s<strong>in</strong>gle hammer.<br />

The grand piano provides an excellent example of a weakly-coupled harmonic oscillator system that has<br />

normal modes. There are either two or three parallel str<strong>in</strong>gs per note that are stretched tightly parallel to the<br />

top of the horizontal sound<strong>in</strong>g board. The str<strong>in</strong>gs press downwards on the bridge that is attached to the top of<br />

the sound<strong>in</strong>g board. The str<strong>in</strong>gs for each note are excited when struck vertically upwards by a s<strong>in</strong>gle hammer.<br />

In the base section of the piano each note comprises two str<strong>in</strong>gs tuned to nearly the same frequency. The<br />

coupl<strong>in</strong>g of the motion of the str<strong>in</strong>gs is via the bridge plus sound<strong>in</strong>g board. Normally, the hammer strikes both<br />

str<strong>in</strong>gs simultaneously excit<strong>in</strong>g the vertical symmetric mode, not the vertical antisymmetric mode. The bridge<br />

is connected to the sound<strong>in</strong>g board which moves the largest amount for the symmetric mode where both str<strong>in</strong>gs<br />

move the bridge <strong>in</strong> phase. This strong coupl<strong>in</strong>g produces a loud sound. The antisymmetric mode does not<br />

move the sound<strong>in</strong>g board much s<strong>in</strong>ce the str<strong>in</strong>gs at the bridge move out of phase. Consequently, the symmetric<br />

mode, that is strongly coupled to the sound<strong>in</strong>g board, damps out more rapidly than the antisymmetric mode<br />

which is weakly coupled to the sound board and thus has a longer time constant for decay s<strong>in</strong>ce the radiated<br />

sound energy is lower than the symmetric mode.<br />

The una-corda pedal (soft pedal) for a grand piano moves the action sideways such that the hammer strikes<br />

only one of the two str<strong>in</strong>gs, or two of the three str<strong>in</strong>gs, result<strong>in</strong>g <strong>in</strong> both the symmetric and antisymmetric<br />

modes be<strong>in</strong>g excited equally. The una-corda pedal produces a characteristically different tone than when the<br />

hammer simultaneously hits the coupled str<strong>in</strong>gs; that is, it produces a smaller transient component. The<br />

symmetric mode rapidly damps due to energy propagation by the sound<strong>in</strong>g board. Thus the longer last<strong>in</strong>g<br />

antisymmetric mode becomes more prom<strong>in</strong>ent when both modes are equally excited us<strong>in</strong>g the una-corda pedal.<br />

The symmetric and antisymmetric modes have slightly different frequencies and produce beats which also<br />

contributes to the different timbre produced us<strong>in</strong>g the una-corda pedal. For the mid and upper frequency<br />

range, the piano has three str<strong>in</strong>gs per note which have onesymmetricmodeandtwoseparateantisymmetric<br />

modes. To further complicate matters, the str<strong>in</strong>gs also can oscillate horizontally which couples weakly to the<br />

bridge plus sound<strong>in</strong>g board. The strengths that these different modes are excited depend on subtle differences<br />

<strong>in</strong> the shape and roughness of the hammer head strik<strong>in</strong>g the str<strong>in</strong>gs. Primarily the hammer excites the two<br />

vertical modes rather than the horizontal modes.


14.6. GENERAL ANALYTIC THEORY FOR COUPLED LINEAR OSCILLATORS 363<br />

14.6 General analytic theory for coupled l<strong>in</strong>ear oscillators<br />

The above discussion of a coupled double-oscillator system has shown that it is possible to select symmetric<br />

and antisymmetric normal modes that are <strong>in</strong>dependent and each have characteristic frequencies. The normal<br />

coord<strong>in</strong>ates for these two normal modes correspond to l<strong>in</strong>ear superpositions of the spatial amplitudes of the<br />

two oscillators and can be obta<strong>in</strong>ed by a rotation <strong>in</strong>to the appropriate normal coord<strong>in</strong>ate system. Extension<br />

of this to systems compris<strong>in</strong>g coupled l<strong>in</strong>ear oscillators, requires development of a general analytic theory,<br />

that is capable of f<strong>in</strong>d<strong>in</strong>g the normal modes plus their eigenvalues and eigenvectors. As illustrated for the<br />

double oscillator, the solution of many coupled l<strong>in</strong>ear oscillators is a classic eigenvalue problem where one has<br />

to rotate to the pr<strong>in</strong>cipal axis system to project out the normal modes. The follow<strong>in</strong>g discussion presents a<br />

general approach to the problem of f<strong>in</strong>d<strong>in</strong>g the normal coord<strong>in</strong>ates for a system of coupled l<strong>in</strong>ear oscillators.<br />

Consider a conservative system of coupled oscillators, described <strong>in</strong> terms of generalized coord<strong>in</strong>ates<br />

and with subscript =1 2 3 for a system with degrees of freedom The coupled oscillators are<br />

assumed to have a stable equilibrium with generalized coord<strong>in</strong>ates 0 at equilibrium. In addition, it is<br />

assumed that the oscillation amplitudes are sufficiently small to ensure that the system is l<strong>in</strong>ear.<br />

For the equilibrium position = 0 the Lagrange equations must satisfy<br />

<br />

˙ = 0 (14.27)<br />

¨ = 0<br />

Every non-zero term of the form ˙ <br />

<strong>in</strong> Lagrange’s equations must conta<strong>in</strong> at least either ˙ or ¨ which<br />

are zero at equilibrium; thus all such terms vanish at equilibrium. That is at equilibrium<br />

µ µ µ <br />

= − =0 (14.28)<br />

<br />

0<br />

where the subscript 0 designates at equilibrium.<br />

14.6.1 K<strong>in</strong>etic energy tensor T<br />

<br />

0<br />

<br />

0<br />

In chapter 76 it was shown that, <strong>in</strong> terms of fixed rectangular coord<strong>in</strong>ates, the k<strong>in</strong>etic energy for bodies,<br />

with generalized coord<strong>in</strong>ates, is expressed as<br />

= 1 2<br />

X<br />

=1 =1<br />

3X<br />

˙ 2 (14.29)<br />

Express<strong>in</strong>g these <strong>in</strong> terms of generalized coord<strong>in</strong>ates = ( ) where =1 2 then the generalized<br />

velocities are given by<br />

X <br />

˙ = ˙ + <br />

(14.30)<br />

<br />

=1 <br />

As discussed <strong>in</strong> chapter 76 if the system is scleronomic then the partial time derivative<br />

<br />

=0 (14.31)<br />

<br />

Thus the k<strong>in</strong>etic energy, equation 1429, of a scleronomic system can be written as a homogeneous quadratic<br />

function of the generalized velocities<br />

= 1 X<br />

˙ ˙ (14.32)<br />

2<br />

where the components of the k<strong>in</strong>etic energy tensor T are<br />

X 3X <br />

≡ <br />

<br />

<br />

<br />

<br />

<br />

<br />

(14.33)<br />

Note that if the velocities ˙ correspond to translational velocity, then the k<strong>in</strong>etic energy tensor T corresponds<br />

to an effective mass tensor, whereas if the velocities correspond to angular rotational velocities, then the<br />

k<strong>in</strong>etic energy tensor T corresponds to the <strong>in</strong>ertia tensor.


364 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

It is possible to make an expansion of the about the equilibrium values of the form<br />

( 1 2 )= ( 0 )+ X <br />

µ<br />

<br />

+ (14.34)<br />

<br />

0<br />

Only the first-order term will be kept s<strong>in</strong>ce the second and higher terms are of the same order as the higherorder<br />

³ terms ignored <strong>in</strong> the Taylor expansion of the potential. Thus, at the equilibrium po<strong>in</strong>t, assume that<br />

<br />

<br />

=0where =1 2 3 .<br />

´0<br />

14.6.2 Potential energy tensor V<br />

Equations 1428 plus 1434 imply that<br />

µ <br />

<br />

0<br />

=0 (14.35)<br />

where =1 2 3 <br />

Make a Taylor expansion about equilibrium for the potential energy, assum<strong>in</strong>g for simplicity that the<br />

coord<strong>in</strong>ates have been translated to ensure that =0at equilibrium. This gives<br />

( 1 2 )= 0 + X <br />

³<br />

<br />

<br />

´0<br />

µ <br />

<br />

0<br />

+ 1 2<br />

X<br />

µ 2 <br />

<br />

+ (14.36)<br />

<br />

0<br />

The l<strong>in</strong>ear term is zero s<strong>in</strong>ce = 0 at the equilibrium po<strong>in</strong>t, and without loss of generality, the<br />

potential can be measured with respect to 0 . Assume that the amplitudes are small, then the expansion<br />

can be restricted to the quadratic term, correspond<strong>in</strong>g to the simple l<strong>in</strong>ear oscillator potential<br />

( 1 2 ) − 0 = 0 ( 1 2 )= 1 X<br />

µ 2 <br />

=<br />

2 <br />

0 1 X<br />

(14.37)<br />

2<br />

That is<br />

<br />

0 ( 1 2 )= 1 X<br />

(14.38)<br />

2<br />

where the components of the potential energy tensor V are def<strong>in</strong>ed as<br />

µ 2 0<br />

≡<br />

<br />

0<br />

Note that the order of differentiation is unimportant and thus the quantity is symmetric<br />

<br />

<br />

(14.39)<br />

= (14.40)<br />

The motion of the system has been specified for small oscillations around the equilibrium position and<br />

it has been shown that 0 ( 1 2 ) has a m<strong>in</strong>imum value at equilibrium which is taken to be zero for<br />

convenience.<br />

In conclusion, equations (1432) and (1438) give<br />

= 1 2<br />

0 = 1 2<br />

X<br />

˙ ˙ (14.41)<br />

<br />

X<br />

(14.42)<br />

where the components of the k<strong>in</strong>etic energy tensor T and potential energy tensor V are<br />

à <br />

!<br />

X 3X <br />

≡ <br />

<br />

<br />

0<br />

µ 2 0<br />

≡<br />

<br />

0<br />

<br />

(14.43)<br />

(14.44)


14.6. GENERAL ANALYTIC THEORY FOR COUPLED LINEAR OSCILLATORS 365<br />

Note that and may have different units, but all the terms <strong>in</strong> the summations for both and 0 have<br />

units of energy. The and values are evaluated at the equilibrium po<strong>in</strong>t, and thus both and <br />

are × arrays of values evaluated at the equilibrium location.<br />

14.6.3 Equations of motion<br />

Both the k<strong>in</strong>etic energy and potential energy terms are products of the coord<strong>in</strong>ates lead<strong>in</strong>g to a set of<br />

coupled equations that are complicated to solve. The problem is greatly simplified by select<strong>in</strong>g a set of<br />

normal coord<strong>in</strong>ates for which both and are diagonal, then the coupl<strong>in</strong>g terms disappear. Thus a<br />

coord<strong>in</strong>ate transformation must be found that simultaneously diagonalizes and <strong>in</strong> order to obta<strong>in</strong> a<br />

set of normal coord<strong>in</strong>ates.<br />

The k<strong>in</strong>etic energy is only a function of generalized velocities while the conservative potential energy<br />

is only a function of the generalized coord<strong>in</strong>ates Thus the Lagrange equations<br />

<br />

− <br />

=0<br />

˙ <br />

(14.45)<br />

reduce to<br />

<br />

+ <br />

=0<br />

˙ <br />

(14.46)<br />

But<br />

<br />

X<br />

= <br />

<br />

<br />

(14.47)<br />

and<br />

<br />

X<br />

= ˙ <br />

˙ <br />

<br />

(14.48)<br />

Thus the Lagrange equations reduce to the follow<strong>in</strong>g set of equations of motion,<br />

X<br />

( + ¨ )=0 (14.49)<br />

<br />

For each where 1 ≤ ≤ there exists a set of second-order l<strong>in</strong>ear homogeneous differential equations<br />

with constant coefficients. S<strong>in</strong>ce the system is oscillatory, it is natural to try a solution of the form<br />

() = (−) (14.50)<br />

Assum<strong>in</strong>g that the system is conservative, then this implies that is real, s<strong>in</strong>ce an imag<strong>in</strong>ary term for <br />

would lead to an exponential damp<strong>in</strong>g term. The arbitrary constants are the real amplitude and the<br />

phase Substitution of this trial solution for each leads to a set of equations<br />

X ¡<br />

− 2 ¢<br />

=0 (14.51)<br />

<br />

where the common factor (−) has been removed. Equation 1451 corresponds to a set of l<strong>in</strong>ear<br />

homogeneous algebraic equations that the amplitudes must satisfy for each . For a non-trivial solution<br />

to exist, the determ<strong>in</strong>ant of the coefficients must vanish, that is<br />

11 − 2 11 12 − 2 12 13 − 2 13 <br />

12 − 2 12 22 − 2 22 23 − 2 23 <br />

13 − 2 13 23 − 2 23 33 − 2 33 <br />

=0 (14.52)<br />

¯ ¯<br />

where the symmetry = has been <strong>in</strong>cluded. This is the standard eigenvalue problem for which<br />

the above determ<strong>in</strong>ant gives the secular equation or the characteristic equation. It is an equation<br />

of degree <strong>in</strong> 2 The roots of this equation are 2 where are the characteristic frequencies or<br />

eigenfrequencies of the normal modes.<br />

Substitution of 2 <strong>in</strong>to equation 1452 determ<strong>in</strong>es the ratio 1 : 2 : 3 : : for this solution<br />

which def<strong>in</strong>es the components of the -dimensional eigenvector a . That is, solution of the secular equations<br />

have determ<strong>in</strong>ed the eigenvalues and eigenvectors of the solutions of the coupled-channel system.


366 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

14.6.4 Superposition<br />

The equations of motion P ( + ¨ )=0are l<strong>in</strong>ear equations that satisfy superposition.<br />

most general solution () canbeasuperpositionofthe eigenvectors a ,thatis<br />

() =<br />

Only the real part of () is mean<strong>in</strong>gful, that is,<br />

() =Re<br />

X<br />

( − )<br />

X<br />

( − ) =<br />

<br />

<br />

Thus the<br />

(14.53)<br />

X<br />

cos ( − ) (14.54)<br />

Thus the most general solution of these l<strong>in</strong>ear equations <strong>in</strong>volves a sum over the eigenvectors of the<br />

system which are cos<strong>in</strong>e functions of the correspond<strong>in</strong>g eigenfrequencies.<br />

14.6.5 Eigenfunction orthonormality<br />

It can be shown that the eigenvectors are orthogonal. In addition, the above procedure only determ<strong>in</strong>es ratios<br />

of amplitudes, thus there is an <strong>in</strong>determ<strong>in</strong>acy that can be used to normalize the . Thus the eigenvectors<br />

form an orthonormal set. Orthonormality of the eigenfunctions for the rank 3 <strong>in</strong>ertia tensor was illustrated<br />

<strong>in</strong> chapter 13102 Similar arguments apply that allow extend<strong>in</strong>g orthonormality to higher rank cases such<br />

that for -body coupled oscillators.<br />

The eigenfunction orthogonality for coupled oscillators can be proved by writ<strong>in</strong>g equation 1451<br />

for both the root and the root. That is,<br />

X<br />

X<br />

= 2 (14.55)<br />

<br />

<br />

X<br />

X<br />

= 2 (14.56)<br />

<br />

Multiply equation 1455 by and sum over . Similarly multiply equation 1456 by and sum over .<br />

These summations lead to<br />

X<br />

X<br />

= 2 (14.57)<br />

<br />

<br />

X<br />

X<br />

= 2 (14.58)<br />

<br />

Note that the left-hand sides of these two equations are identical. Thus tak<strong>in</strong>g the difference between these<br />

equations gives<br />

¡ X<br />

<br />

2<br />

− 2 ¢<br />

=0 (14.59)<br />

<br />

Note that if ¡ 2 − ¢ 2 6=0, that is, assum<strong>in</strong>g that the eigenfrequencies are not degenerate, then to ensure<br />

that equation 1459 is zero requires that<br />

X<br />

=0 6= (14.60)<br />

<br />

This shows that the eigenfunctions are orthogonal. If the eigenfrequencies are degenerate, i.e. 2 = 2 ,<br />

then, with no loss of generality, the axes and can be chosen to be orthogonal.<br />

The eigenfunction normalization can be chosen freely s<strong>in</strong>ce only ratios of the eigenfunction components<br />

are determ<strong>in</strong>ed when is used <strong>in</strong> equation 1451. The k<strong>in</strong>etic energy, given by equation 1432<br />

must be positive, or zero for the case of a static system. That is<br />

= 1 2<br />

<br />

<br />

<br />

X<br />

˙ ˙ ≥ 0 (14.61)


14.6. GENERAL ANALYTIC THEORY FOR COUPLED LINEAR OSCILLATORS 367<br />

Use the time derivative of equation 1454 to determ<strong>in</strong>e ˙ and <strong>in</strong>sert <strong>in</strong>to equation 1461 gives that the k<strong>in</strong>etic<br />

energy is<br />

= 1 X<br />

˙ ˙ = 1 X X<br />

cos ( − ) cos ( − ) (14.62)<br />

2<br />

2<br />

<br />

<br />

<br />

For the diagonal term = <br />

"<br />

#<br />

= 1 X<br />

1<br />

X<br />

X<br />

˙ ˙ = 2 cos 2 ( − ) ≥ 0 (14.63)<br />

2<br />

2<br />

<br />

<br />

<br />

S<strong>in</strong>cetheterm<strong>in</strong>thesquarebracketsmustbepositive,then<br />

X<br />

≥ 0 (14.64)<br />

<br />

S<strong>in</strong>ce this sum must be a positive number, and the magnitude of the amplitudes can be chosen freely, then<br />

it is possible to normalize the eigenfunction amplitudes to unity. That is, choose that<br />

X<br />

=1 (14.65)<br />

<br />

The orthogonality equation, 1460 and the normalization equation 1465 can be comb<strong>in</strong>ed <strong>in</strong>to a s<strong>in</strong>gle<br />

orthonormalization equation<br />

X<br />

= (14.66)<br />

<br />

This has shown that the eigenvectors form an orthonormal set.<br />

S<strong>in</strong>ce the component of the eigenvector is ,thenthe eigenvector can be written <strong>in</strong> the form<br />

a = X <br />

be (14.67)<br />

where be are the unit vectors for the generalized coord<strong>in</strong>ates.<br />

14.6.6 Normal coord<strong>in</strong>ates<br />

The above general solution of the coupled-oscillator problem is best expressed <strong>in</strong> terms of the normal coord<strong>in</strong>ates<br />

which are <strong>in</strong>dependent. It is more transparent if the superposition of the normal modes are written<br />

<strong>in</strong> the form<br />

X<br />

() = <br />

(14.68)<br />

<br />

where the complex factor <strong>in</strong>cludes the arbitrary scale factor to allow for arbitrary amplitudes as well<br />

as the fact that the amplitudes have been normalized and the phase factor has been chosen.<br />

Def<strong>in</strong>e<br />

() ≡ <br />

(14.69)<br />

then equation 1468 can be written as<br />

X<br />

() = () (14.70)<br />

Equation 1470 can be expressed schematically as the matrix multiplication<br />

The () are the normal coord<strong>in</strong>ates whichcanbeexpressed<strong>in</strong>theform<br />

<br />

q = {a}·η (14.71)<br />

η = {a} −1 q (14.72)<br />

Each normal mode corresponds to a s<strong>in</strong>gle eigenfrequency, which satisfies the l<strong>in</strong>ear oscillator equation<br />

¨ + 2 =0 (14.73)


368 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

14.7 Two-body coupled oscillator systems<br />

The two-body coupled oscillator is the simplest coupled-oscillator system that illustrates the general features<br />

of coupled oscillators. The follow<strong>in</strong>g four examples <strong>in</strong>volve parallel and series coupl<strong>in</strong>gs of two l<strong>in</strong>ear<br />

oscillators or two plane pendula.<br />

14.2 Example: Two coupled l<strong>in</strong>ear oscillators<br />

The coupled double-oscillator problem, figure 141 discussed <strong>in</strong> chapter 142, can be used to demonstrate<br />

that the general analytic theory gives the same solution as obta<strong>in</strong>ed by direct solution of the equations of<br />

motion <strong>in</strong> chapter 142.<br />

1) The first stage is to determ<strong>in</strong>e the potential and k<strong>in</strong>etic energies us<strong>in</strong>g an appropriate set of generalized<br />

coord<strong>in</strong>ates, which here are 1 and 2 .Thepotentialenergyis<br />

= 1 2 2 1 + 1 2 2 2 + 1 2 0 ( 2 − 1 ) 2 = 1 2 ( + 0 ) 2 1 + 1 2 ( + 0 ) 2 2 − 0 1 2<br />

whilethek<strong>in</strong>eticenergyisgivenby<br />

= 1 2 ˙2 1 + 1 2 ˙2 2<br />

2) The second stage is to evaluate the potential energy and k<strong>in</strong>etic energy tensors. The potential<br />

energy tensor is nondiagonal s<strong>in</strong>ce gives<br />

11<br />

≡<br />

12 =<br />

µ 2 <br />

= + <br />

1 1<br />

0<br />

0 = 22<br />

µ 2 <br />

= −<br />

1 2<br />

0<br />

0 = 21<br />

That is, the potential energy tensor is<br />

½ ¾<br />

+ <br />

0<br />

−<br />

V =<br />

0<br />

− 0 + 0<br />

Similarly, the k<strong>in</strong>etic energy is given by<br />

= 1 2 ˙2 1 + 1 2 ˙2 2 = 1 X<br />

˙ ˙ <br />

2<br />

S<strong>in</strong>ce 11 = 22 = and 12 = 21 =0then the k<strong>in</strong>etic energy tensor is<br />

½ ¾ 0<br />

T =<br />

0 <br />

Note that for this case, the k<strong>in</strong>etic energy tensor equals the mass tensor, which is diagonal, whereas the<br />

potential energy tensor equals the spr<strong>in</strong>g constant tensor, which is nondiagonal.<br />

3) Thethirdstageistousethepotentialenergy and k<strong>in</strong>etic energy tensors to evaluate the secular<br />

determ<strong>in</strong>ant us<strong>in</strong>g equations 1452<br />

¯ + 0 − 2 − 0<br />

− 0 + 0 − <br />

¯¯¯¯ 2 =0<br />

The expansion of this secular determ<strong>in</strong>ant yields<br />

¡<br />

+ 0 − 2¢ 2<br />

− 02 =0<br />

<br />

That is<br />

¡<br />

+ 0 − 2¢ = ± 0


14.7. TWO-BODY COUPLED OSCILLATOR SYSTEMS 369<br />

Solv<strong>in</strong>g for gives<br />

The solutions are<br />

r<br />

+ 0 ± <br />

=<br />

0<br />

<br />

1 =<br />

2 =<br />

r<br />

+2<br />

0<br />

r <br />

<br />

<br />

which is the same as derived previously, (equations 147 − 9).<br />

4) The fourth step is to <strong>in</strong>sert either one of these eigenfrequencies <strong>in</strong>to the secular equation<br />

X ¡<br />

− 2 ¢<br />

=0 ()<br />

Consider the secular equation for =1<br />

¡<br />

+ 0 − 2 ¢ 1 − 0 2 =0<br />

<br />

Then for the first eigenfrequency 1 that is, =1 =1<br />

which simplifies to<br />

( + 0 − − 2 0 ) 11 − 0 21 =0<br />

= 11 = − 21<br />

Similarly, for the other eigenfrequency 2 , that is, =1 =2<br />

which simplifies to<br />

( + 0 − ) 12 − 0 22 =0<br />

= 12 = 22<br />

5) The f<strong>in</strong>al stage is to write the general coord<strong>in</strong>ates <strong>in</strong> terms of the normal coord<strong>in</strong>ates () ≡<br />

Thus<br />

1 = 11 1 + 12 2 = 11 1 + 22 2<br />

and<br />

2 = 21 1 + 22 2 = − 11 1 + 22 2<br />

Add<strong>in</strong>g or subtract<strong>in</strong>g gives that the normal modes are<br />

1 =<br />

1<br />

( 1 − 2 )<br />

2 11<br />

2 =<br />

1<br />

( 2 + 1 )<br />

2 22<br />

Thus the symmetric normal mode 2 corresponds to an oscillation of the center-of-mass with the lower<br />

frequency 2 = p <br />

This frequency is the same as for one s<strong>in</strong>gle mass on a spr<strong>in</strong>g of spr<strong>in</strong>g constant<br />

which is as expected s<strong>in</strong>ce they vibrate <strong>in</strong> unison andqthus the coupl<strong>in</strong>g spr<strong>in</strong>g force does not act. The<br />

+2<br />

antisymmetric mode 1 has the higher frequency 1 =<br />

<br />

s<strong>in</strong>ce the restor<strong>in</strong>g force <strong>in</strong>cludes both the<br />

ma<strong>in</strong> spr<strong>in</strong>g plus the coupl<strong>in</strong>g spr<strong>in</strong>g.<br />

The above example illustrates that the general analytic theory for coupled l<strong>in</strong>ear oscillators gives the<br />

same answer as obta<strong>in</strong>ed <strong>in</strong> chapter 142 us<strong>in</strong>g Newton’s equations of motion. However, the general analytic<br />

theory is a more powerful technique for solv<strong>in</strong>g complicated coupled oscillator systems. Thus the general<br />

analytic theory will be used for solv<strong>in</strong>g all the follow<strong>in</strong>g coupled oscillator problems.


370 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

14.3 Example: Two equal masses series-coupled by two equal spr<strong>in</strong>gs<br />

Consider the series-coupled system shown <strong>in</strong> the figure.<br />

1) The first stage is to determ<strong>in</strong>e the potential and k<strong>in</strong>etic<br />

energies us<strong>in</strong>g an appropriate set of generalized coord<strong>in</strong>ates,<br />

which here are 1 and 2 .Thepotentialenergyis<br />

1 2<br />

= 1 2 2 1 + 1 2 ( 2 − 1 ) 2 = 2 1 + 1 2 2 2 − 1 2<br />

whilethek<strong>in</strong>eticenergyisgivenby<br />

= 1 2 ˙2 1 + 1 2 ˙2 2<br />

Two equal masses series-coupled by two<br />

equal spr<strong>in</strong>gs.<br />

2) The second stage is to evaluate the potential energy and mass tensors. The potential energy tensor<br />

is nondiagonal s<strong>in</strong>ce gives<br />

11<br />

≡<br />

12 =<br />

22 =<br />

µ 2 <br />

=2<br />

1 1<br />

0<br />

µ 2 <br />

= − = 21<br />

1 2<br />

0<br />

µ 2 <br />

= <br />

2 2<br />

0<br />

That is, the potential energy tensor is<br />

½ 2<br />

V =<br />

−<br />

−<br />

<br />

¾<br />

Similarly, s<strong>in</strong>ce the k<strong>in</strong>etic energy is given by<br />

= 1 2 ˙2 1 + 1 2 ˙2 2 = 1 X<br />

˙ ˙ <br />

2<br />

then 11 = 22 = and 12 = 21 =0 Thus the k<strong>in</strong>etic energy tensor is<br />

½ ¾ 0<br />

T =<br />

0 <br />

Note that for this case the k<strong>in</strong>etic energy tensor is diagonal whereas the potential energy tensor is nondiagonal.<br />

3) Thethirdstageistousethepotentialenergy and k<strong>in</strong>etic energy tensors to evaluate the secular<br />

determ<strong>in</strong>ant us<strong>in</strong>g equation 1452<br />

2 − <br />

¯¯¯¯ 2 −<br />

− − <br />

¯¯¯¯ 2 =0<br />

The expansion of this secular determ<strong>in</strong>ant yields<br />

¡<br />

2 − <br />

2 ¢¡ − 2¢ − 2 =0<br />

<br />

That is<br />

4 − 3 2 + 2<br />

2 =0<br />

The solutions are<br />

√ r √ r 5+1 5 − 1 <br />

1 =<br />

2 =<br />

2 <br />

2 <br />

4) The fourth step is to <strong>in</strong>sert these eigenfrequencies <strong>in</strong>to the secular equation 1451<br />

X ¡<br />

− 2 ¢<br />

=0


14.7. TWO-BODY COUPLED OSCILLATOR SYSTEMS 371<br />

Consider =1<strong>in</strong> the above equation<br />

¡<br />

2 − <br />

2<br />

¢ 1 − 2 =0<br />

Then for eigenfrequency 1 ,thatis, =1=1<br />

√<br />

5 − 1<br />

11 = − 21<br />

2<br />

Similarly, for =1 =2 √<br />

5+1<br />

12 = 22<br />

2<br />

5) The f<strong>in</strong>al stage is to write the general coord<strong>in</strong>ates <strong>in</strong> terms of the normal coord<strong>in</strong>ates () ≡<br />

<br />

Thus<br />

1 = 11 1 + 12 2 = 11 1 + 2 22<br />

√ 2 5+1<br />

and<br />

Ã√ !<br />

5 − 1<br />

2 = 21 1 + 22 2 = −<br />

11 <br />

2<br />

1 + 22 2<br />

Add<strong>in</strong>g or subtract<strong>in</strong>g gives that the normal modes are<br />

à Ã√ ! !<br />

1<br />

5 − 1<br />

1 = √ 1 −<br />

2<br />

11 5 2<br />

à Ã√ ! !<br />

1<br />

5+1<br />

2 = √ 1 +<br />

2<br />

22 5 2<br />

Thusthesymmetricnormalmodehasthelowerfrequency 2 = √ p<br />

5−1 <br />

2 The antisymmetric mode has the<br />

frequency 1 = √ p<br />

5+1 <br />

2 <br />

s<strong>in</strong>ce both spr<strong>in</strong>gs provide the restor<strong>in</strong>g force. This case is <strong>in</strong>terest<strong>in</strong>g <strong>in</strong> that for<br />

both normal modes, the amplitudes for the motion of the two masses are different.<br />

14.4 Example: Two parallel-coupled plane pendula<br />

Consider the coupled double pendulum system shown <strong>in</strong><br />

the adjacent figure, which comprises two parallel plane pendula<br />

weakly coupled by a spr<strong>in</strong>g. The angles 1 and 2 are<br />

chosen to be the generalized coord<strong>in</strong>ates and the potential energy<br />

is chosen to be zero at equilibrium. Then the k<strong>in</strong>etic<br />

energy is<br />

= 1 ³<br />

2 1<br />

³<br />

˙ 1´2<br />

+<br />

2 ˙ 2´2<br />

As discussed <strong>in</strong> chapter 3, it is necessary to make the smallangle<br />

approximation <strong>in</strong> order to make the equations of motion<br />

for the simple pendulum l<strong>in</strong>ear and solvable analytically. That<br />

is,<br />

1 2<br />

Two parallel-coupled plane pendula.<br />

= (1 − cos 1 )+ (1 − cos 2 )+ 1 2 ( s<strong>in</strong> 1 − s<strong>in</strong> 2 ) 2<br />

k<br />

'<br />

<br />

2<br />

¡<br />

<br />

2<br />

1 + 2 2¢<br />

+<br />

2<br />

2 ( 1 − 2 ) 2<br />

assum<strong>in</strong>g the small angle approximation s<strong>in</strong> ≈ and (1 − cos 1 )= 2<br />

2 <br />

Thesecondstageistoevaluatethek<strong>in</strong>eticenergy and potential energy tensors<br />

½ ¾<br />

½ ¾<br />

<br />

2<br />

0<br />

+ <br />

2<br />

−<br />

T =<br />

0 2 V =<br />

2<br />

− 2 + 2


372 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

Note that for this case the k<strong>in</strong>etic energy tensor is diagonal whereas the potential energy tensor is nondiagonal.<br />

The third stage is to evaluate the secular determ<strong>in</strong>ant<br />

¯ + 2 − 2 2 − 2<br />

− 2 + 2 − 2 <br />

¯¯¯¯ 2 =0<br />

which gives the characteristic equation<br />

¡<br />

+ 2 − 2 2¢ 2 ¡<br />

= <br />

2 ¢ 2<br />

or<br />

The two solutions are<br />

+ − 2 = ±<br />

2 1 = 2 2 = <br />

+ 2 <br />

The fourth step is to <strong>in</strong>sert these eigenfrequencies <strong>in</strong>to equation 1451<br />

X ¡<br />

− 2 ¢<br />

=0<br />

<br />

Consider =1 ¡<br />

+ 2 − 2 2¢ 1 − 2 2 =0<br />

Then for the first eigenfrequency, 1 , the subscripts are =1 =1<br />

³ + 2 − 2´<br />

11 − 2 21 =0<br />

which simplifies to<br />

Similarly, for =1 =2<br />

11 = 21<br />

µ<br />

µ <br />

+ 2 −<br />

+ 2 <br />

2 12 − 2 22 =0<br />

<br />

which simplifies to<br />

12 = − 22<br />

The f<strong>in</strong>al stage is to write the general coord<strong>in</strong>ates <strong>in</strong> terms of the normal coord<strong>in</strong>ates<br />

and<br />

1 = 11 1 + 12 2 = 11 1 − 22 2<br />

2 = 21 1 + 22 2 = 11 1 + 22 2<br />

Add<strong>in</strong>g or subtract<strong>in</strong>g these equations gives that the normal modes are<br />

1 = 1 ( 1 + 2 ) <br />

2 2 = 1 ( 2 − 1 )<br />

11 2 22<br />

As for the case of the double oscillator discussed <strong>in</strong> example 142, the symmetric normal mode corresponds<br />

to an oscillation of the center-of-mass, with zero relative motion of the two pendula, which has the lower<br />

frequency 1 = p <br />

<br />

This frequency is the same as for one <strong>in</strong>dependent pendulum as expected s<strong>in</strong>ce they<br />

vibrate <strong>in</strong> unison and thus the only restor<strong>in</strong>g force is gravity. The antisymmetric mode corresponds q to ¡<br />

relative motion of the two pendula with stationary center-of-mass and has the frequency 2 = <br />

+ ¢<br />

2<br />

<br />

s<strong>in</strong>ce the restor<strong>in</strong>g force <strong>in</strong>cludes both the coupl<strong>in</strong>g spr<strong>in</strong>g and gravity.<br />

This example <strong>in</strong>troduces the role of degeneracy which occurs <strong>in</strong> this system if the coupl<strong>in</strong>g of the pendula<br />

is zero, that is, =0 lead<strong>in</strong>g to both frequencies be<strong>in</strong>g equal, i.e. 1 = 2 = p <br />

<br />

. When =0,thenboth<br />

{T} and {V} are diagonal and thus <strong>in</strong> the ( 1 2 ) space the two pendula are <strong>in</strong>dependent normal modes.<br />

However, the symmetric and asymmetric normal modes, as derived above, are equally good normal modes.<br />

In fact, s<strong>in</strong>ce the modes are degenerate, any l<strong>in</strong>ear comb<strong>in</strong>ation of the motion of the <strong>in</strong>dependent pendula are<br />

equally good normal modes and thus one can use any set of orthogonal normal modes to describe the motion.


14.7. TWO-BODY COUPLED OSCILLATOR SYSTEMS 373<br />

14.5 Example: The series-coupled double plane pendula<br />

The double-pendula system comprises one plane pendulum attached<br />

to the end of another plane pendulum both oscillat<strong>in</strong>g <strong>in</strong> the same plane.<br />

The k<strong>in</strong>etic and potential energies for this system are given <strong>in</strong> example<br />

621 to be<br />

= 1 2 ( 1 + 2 ) 2 ˙ 2 1 1 + 2 1 2 ˙1 ˙2 cos( 1 − 2 )+ 1 2 2 2 ˙ 2 2 2<br />

= ( 1 + 2 ) 1 (1 − cos 1 )+ 2 2 (1 − cos 2 )<br />

a) Small-amplitude l<strong>in</strong>ear regime<br />

Use of the small-angle approximation makes this system l<strong>in</strong>ear and<br />

solvable analytically. That is, and become<br />

Two series-coupled plane pendula.<br />

= 1 2 ( 1 + 2 ) 1 2 1 + 1 2 2 2 2 2<br />

= 1 2 ( 1 + 2 ) 2 1 ˙ 2 1 + 2 1 2 ˙1 ˙2 + 1 2 2 2 2 ˙ 2 2<br />

Thus the k<strong>in</strong>etic energy and potential energy tensors are<br />

½ ¾<br />

½ ¾<br />

(1 + <br />

T =<br />

2 ) 2 1 2 1 2<br />

(1 + <br />

2 1 2 2 2 V =<br />

2 ) 1 0<br />

2<br />

0 2 2<br />

Note that T is nondiagonal, whereas V is diagonal which is opposite<br />

to the case of the two parallel-coupled plane pendula.<br />

The solution of this case is simpler if it is assumed that 1 = 2 = <br />

and 1 = 2 = . Then<br />

½ ¾<br />

½ ¾<br />

T = 2 2 1<br />

V = <br />

1 1<br />

2 2<br />

2<br />

0 0<br />

0 2 0<br />

where 0 = p <br />

<br />

which is the frequency of a s<strong>in</strong>gle pendulum.<br />

The next stage is to evaluate the secular determ<strong>in</strong>ant<br />

<br />

¯¯¯¯ 2 2( 2 0 − 2 ) − 2<br />

− 2 ( 2 0 − 2 ) ¯ =0<br />

The eigenvalues are<br />

2 1 =(2− √ 2) 2 0 2 2 =(2+ √ 2) 2 0<br />

As shown <strong>in</strong> the adjacent figure, the normal modes for this system<br />

are<br />

1 = 1 (<br />

2 1 + 2<br />

√ ) 2 = 1 (<br />

11 2 2 1 − 2<br />

√ )<br />

22 2<br />

Normal modes for two<br />

series-coupled plane pendula.<br />

The second mass has a √ 2 larger amplitude that is <strong>in</strong> phase for solution 1 and out of phase for solution 2.<br />

b) Large amplitude chaotic regime<br />

Stachowiak and Okada [Sta05] used computer simulations to numerically analyze the behavior of this<br />

system with <strong>in</strong>crease <strong>in</strong> the oscillation amplitudes. Po<strong>in</strong>caré sections, bifurcation diagrams, and Lyapunov<br />

exponents all confirm that this system evolves from regular normal-mode oscillatory behavior <strong>in</strong> the l<strong>in</strong>ear<br />

regime at low energy, to chaotic behavior at high excitation energies where non-l<strong>in</strong>earity dom<strong>in</strong>ates. This<br />

behavior is analogous to that of the driven, l<strong>in</strong>early-damped, harmonic pendulum described <strong>in</strong> chapter 35


374 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

14.8 Three-body coupled l<strong>in</strong>ear oscillator systems<br />

Chapter 147 discussed parallel and series arrangements of two coupled oscillators. Extend<strong>in</strong>g from two to<br />

three coupled l<strong>in</strong>ear oscillators <strong>in</strong>troduces <strong>in</strong>terest<strong>in</strong>g new characteristics of coupled oscillator systems. For<br />

more than two coupled oscillators, coupled oscillator systems separate <strong>in</strong>to two classifications depend<strong>in</strong>g on<br />

whether each oscillator is coupled to the rema<strong>in</strong><strong>in</strong>g − 1 oscillators, or when the coupl<strong>in</strong>g is only to the<br />

nearest neighbors as illustrated below.<br />

14.6 Example: Three plane pendula; mean-field l<strong>in</strong>ear coupl<strong>in</strong>g<br />

Consider three identical pendula with mass m and length<br />

, suspended from a common support that yields slightly to<br />

pendulum motion lead<strong>in</strong>g to a coupl<strong>in</strong>g between all three pendula<br />

as illustrated <strong>in</strong> the adjacent figure. Assume that the<br />

motion of the three pendula all are <strong>in</strong> the same plane. This<br />

case is analogous to the piano where three str<strong>in</strong>gs <strong>in</strong> the treble<br />

section are coupled by the slightly-yield<strong>in</strong>g common bridge<br />

plus sound<strong>in</strong>g board lead<strong>in</strong>g to coupl<strong>in</strong>g between each of the<br />

three coupled oscillators. This case illustrates the important<br />

concept of degeneracy.<br />

The generalized coord<strong>in</strong>ates are the angles 1 2 and 3 <br />

Assume that the support yields such that the actual deflection<br />

angle for pendulum 1 is<br />

0 1 = 1 − 2 ( 2 + 3 )<br />

b b b<br />

m m m<br />

Three plane pendula with complete l<strong>in</strong>ear<br />

coupl<strong>in</strong>g.<br />

where the coupl<strong>in</strong>g coefficient is small and <strong>in</strong>volves all the pendula, not just the nearest neighbors. Assume<br />

that the same coupl<strong>in</strong>g relation exists for the other angle coord<strong>in</strong>ates. The gravitational potential energy of<br />

each pendulum is given by<br />

1 = (1 − cos 1 ) ≈ 1 2 2 1<br />

assum<strong>in</strong>g the small angle approximation. Ignor<strong>in</strong>g terms of order 2 gives that the potential energy<br />

= <br />

2<br />

¡<br />

<br />

02<br />

1 + 02<br />

2 + 3<br />

02 ¢ ¡<br />

= <br />

2<br />

2 1 + 2 2 + 2 ¢<br />

3 − 2 1 2 − 2 1 3 − 2 2 3<br />

The k<strong>in</strong>etic energy evaluated at the equilibrium location is<br />

= 1 2 ³<br />

˙ 1´2<br />

+<br />

1<br />

2 ³<br />

˙ 2´2<br />

+<br />

1<br />

2 ³<br />

˙ 3´2<br />

The next stage is to evaluate the {T} and {V} tensors<br />

⎧ ⎫<br />

⎨ 1 0 0 ⎬<br />

T = 2 0 1 0<br />

⎩ ⎭<br />

0 0 1<br />

⎧<br />

⎨<br />

V = <br />

⎩<br />

1 − −<br />

− 1 −<br />

− − 1<br />

The third stage is to evaluate the secular determ<strong>in</strong>ant which can be written as<br />

1 − 2 − −<br />

<br />

− 1 − 2 − =0<br />

¯ − − 1 −<br />

¯¯¯¯¯¯¯<br />

2<br />

Expand<strong>in</strong>g and factor<strong>in</strong>g gives<br />

µ µ µ <br />

2 − 1 − <br />

2 − 1 − <br />

2 − 1+2<br />

=0<br />

⎫<br />

⎬<br />


14.8. THREE-BODY COUPLED LINEAR OSCILLATOR SYSTEMS 375<br />

The roots are<br />

r r r √ √ √<br />

1 = 1+ 2 = 1+ 3 = 1 − 2<br />

<br />

<br />

<br />

This case results <strong>in</strong> two degenerate eigenfrequencies, 1 = 2 while 3 is the lowest eigenfrequency.<br />

The eigenvectors can be determ<strong>in</strong>ed by substitution of the eigenfrequencies <strong>in</strong>to<br />

X ¡<br />

− 2 ¢<br />

=0<br />

<br />

Consider the lowest eigenfrequency 3 i.e. =3 for =1 and substitute for 3 = p <br />

<br />

√ 1 − 2 gives<br />

while for =3 =2<br />

Solv<strong>in</strong>g these gives<br />

2 13 − 23 − 33 =0<br />

− 13 +2 23 − 33 =0<br />

13 = 23 = 33<br />

Assum<strong>in</strong>g that the eigenfunction is normalized to unity<br />

2 13 + 2 23 + 2 33 =1<br />

then for the third eigenvector 3<br />

13 = 23 = 33 = 1 √<br />

3<br />

This solution corresponds to all three pendula oscillat<strong>in</strong>g <strong>in</strong> phase with the same amplitude, that is, a coherent<br />

oscillation.<br />

Derivation of the eigenfunctions for the other two eigenfrequencies is complicated because of the degeneracy<br />

1 = 2 there are only five <strong>in</strong>dependent equations to specify the six unknowns for the eigenvectors<br />

1 and 2 That is, the eigenvectors can be chosen freely as long as the orthogonality and normalization are<br />

satisfied. For example, sett<strong>in</strong>g 31 =0 to remove the <strong>in</strong>determ<strong>in</strong>acy, results <strong>in</strong> the a matrix<br />

⎧ √ √ ⎫<br />

1<br />

⎨ 2 2<br />

1<br />

6 6<br />

1<br />

√ 3√<br />

3 ⎬<br />

{a} = − 1<br />

⎩ 2√<br />

2<br />

1<br />

6 6<br />

1<br />

3√<br />

√ 3<br />

0 − 3√ 1 6<br />

1 ⎭<br />

3 3<br />

and thus the solution is given by<br />

⎧<br />

⎨ 1<br />

2<br />

⎩<br />

3<br />

⎫<br />

⎬<br />

⎭ = ⎧<br />

⎨<br />

⎩<br />

√ √ ⎫ ⎧<br />

1<br />

2 2<br />

1<br />

6 6<br />

1<br />

√ 3√<br />

3 ⎬ ⎨ 1<br />

− 2√ 1 2<br />

1<br />

6 6<br />

1<br />

3√<br />

√ 3 <br />

0 − 3√ 1 6<br />

1 ⎭ ⎩ 2<br />

3 3 3<br />

The normal modes are obta<strong>in</strong>ed by tak<strong>in</strong>g the <strong>in</strong>verse matrix {a} −1 and us<strong>in</strong>g {η} = {a} −1 {θ} Note<br />

that s<strong>in</strong>ce {a} is real and orthogonal, then {a} −1 equals the transpose of {a} That is;<br />

⎧ ⎫ ⎧ √ ⎫ ⎧ ⎫<br />

⎨ 1 ⎬ ⎨<br />

1<br />

<br />

⎩ 2<br />

⎭ = 2 2 −<br />

1<br />

√ 2√<br />

√ 2 0 ⎬ ⎨ 1 ⎬<br />

1<br />

⎩<br />

6 6<br />

1<br />

6 6 −<br />

1<br />

√ √ 3√<br />

√ 6 2<br />

<br />

1<br />

3<br />

3 3<br />

1<br />

3 3<br />

1 ⎭ ⎩ ⎭<br />

3 3 3<br />

The normal mode 3 has eigenfrequency<br />

r √<br />

3 = 1 − 2<br />

<br />

⎫<br />

⎬<br />

⎭<br />

and eigenvector<br />

η 3 = 1 √<br />

3<br />

( 1 2 3 )


376 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

This corresponds to the <strong>in</strong>-phase oscillation of all three pendula.<br />

The other two degenerate solutions are<br />

η 1 = 1 √<br />

2<br />

( 1 − 2 0) η 2 = 1 √<br />

6<br />

( 1 2 −2 3 )<br />

with eigenvalues<br />

r √<br />

1 = 2 = 1+<br />

<br />

These two degenerate normal modes correspond to two pendula oscillat<strong>in</strong>g out of phase with the same amplitude,<br />

or two oscillat<strong>in</strong>g <strong>in</strong> phase with the same amplitude and the third out of phase with twice the amplitude.<br />

An important result of this toy model is that the most symmetric mode 3 is pushed far from all the other<br />

modes. Note that for this example, the coherent mode 3 corresponds to the center-of-mass oscillation with<br />

no relative motion between the three pendula. This is <strong>in</strong> contrast to the eigenvectors 1 and 2 which both<br />

correspond to relative motion of the pendula such that there is zero center-of-mass motion. This mean-field<br />

coupl<strong>in</strong>g behavior is exhibited by collective motion <strong>in</strong> nuclei as discussed <strong>in</strong> example 1412.<br />

14.7 Example: Three plane pendula; nearest-neighbor coupl<strong>in</strong>g<br />

There is a large and important class of coupled oscillators<br />

where the coupl<strong>in</strong>g is only between nearest neighbors; a crystall<strong>in</strong>e<br />

lattice is a classic example. A toy model for such a<br />

system is the case of three identical pendula coupled by two<br />

identical spr<strong>in</strong>gs, where only the nearest neighbors are coupled<br />

as shown <strong>in</strong> the adjacent figure. Assume the identical<br />

pendula are of length and mass . As <strong>in</strong> the last example,<br />

the k<strong>in</strong>etic energy evaluated at the equilibrium location is<br />

= 1 2 2 ˙2 1 + 1 2 2 ˙2 2 + 1 2 2 ˙2 3<br />

The gravitational potential energy of each pendulum equals<br />

(1 − cos ) ≈ 1 2 2 thus<br />

while the potential energy <strong>in</strong> the spr<strong>in</strong>gs is given by<br />

= 1 2 (2 1 + 2 2 + 2 3)<br />

1 2 3<br />

Three plane pendula with nearest-neighbour<br />

coupl<strong>in</strong>g.<br />

= 1 2 2 h ( 2 − 1 ) 2 +( 3 − 2 ) 2i = 1 2 2 £ 2 1 +2 2 2 + 2 3 − 2 1 2 − 2 2 3<br />

¤<br />

Thus the total potential energy is given by<br />

= 1 2 (2 1 + 2 2 + 2 3 )+1 2 2 £ 2 1 +22 2 + 2 3 − 2 1 2 − 2 2 3<br />

¤<br />

The Lagrangian then becomes<br />

= 1 2<br />

³˙2 1 +<br />

2 ˙ 2 2 + ˙<br />

´<br />

2<br />

3 − 1 ¡<br />

+ <br />

2 ¢ 2 1 + 1 ¡<br />

+2<br />

2 ¢ 2 2 + 1 ¡<br />

+ <br />

2 ¢ 2 3 − 2 ( 1 2 + 2 3 )<br />

2<br />

2<br />

2<br />

Us<strong>in</strong>g this <strong>in</strong> the Euler-Lagrange equations gives the equations of motion<br />

T = 2 ⎧<br />

⎨<br />

⎩<br />

1 0 0<br />

0 1 0<br />

0 0 1<br />

2¨1 − ( + 2 ) 1 + 2 2 = 0<br />

2¨2 − ( +2 2 ) 2 + 2 ( 1 + 3 ) = 0<br />

2¨3 − ( + 2 ) 3 + 2 2 = 0<br />

The general analytic approach requires the and energy tensors given by<br />

⎫<br />

⎧<br />

⎬<br />

⎨<br />

⎭<br />

V =<br />

⎩<br />

+ 2 − 2 0<br />

− 2 +2 2 − 2<br />

0 − 2 + 2 ⎫<br />

⎬<br />


14.8. THREE-BODY COUPLED LINEAR OSCILLATOR SYSTEMS 377<br />

Note that <strong>in</strong> contrast to the prior case of three fully-coupled pendula, for the nearest neighbor case the potential<br />

energy tensor {V} is non-zero only on the diagonal and ±1 components parallel to the diagonal.<br />

The third stage is to evaluate the secular determ<strong>in</strong>ant of the ¡ V − 2 T ¢ matrix, that is<br />

¯<br />

+ 2 − 2 2 − 2 0<br />

− 2 +2 2 − 2 2 − 2<br />

0 − 2 + 2 − 2 2 ¯¯¯¯¯¯<br />

=0<br />

This results <strong>in</strong> the characteristic equation<br />

¡<br />

− 2 2¢¡ + 2 − 2 2¢¡ +3 2 − 2 2¢ =0<br />

which results <strong>in</strong> the three non-degenerate eigenfrequencies for the normal modes.<br />

The normal modes are similar to the prior case of complete l<strong>in</strong>ear<br />

coupl<strong>in</strong>g, as shown <strong>in</strong> the adjacent figure.<br />

1 = p <br />

<br />

This lowest mode 1 <strong>in</strong>volves the three pendula oscillat<strong>in</strong>g<br />

<strong>in</strong> phase such that the spr<strong>in</strong>gs are not stretched or compressed thus the<br />

period of this coherent oscillation is the same as an <strong>in</strong>dependent pendulum<br />

of mass and length . That is<br />

η 1 = √ 1 ( 1 2 3 )<br />

3<br />

1<br />

2 = p <br />

+ This second mode 2 has the central mass stationary with<br />

the outer pendula oscillat<strong>in</strong>g with the same amplitude and out of phase.<br />

That is<br />

η 2 = 1 √<br />

2<br />

( 1 0 − 3 )<br />

q<br />

<br />

3 =<br />

+ 3 . This third mode 3 <strong>in</strong>volves the outer pendula <strong>in</strong> phase<br />

with the same amplitude while the central pendulum oscillat<strong>in</strong>g with angle<br />

3 = −2 1 .Thatis<br />

η 3 = √ 1 ( 1 −2 2 3 )<br />

6<br />

2<br />

Similar to the prior case of three completely-coupled pendula, the coherent<br />

normal mode η 1 corresponds to an oscillation of the center-of-mass with<br />

no relative motion, while η 2 and η 3 correspond to relative motion of<br />

the pendula with stationary center of mass motion. In contrast to the<br />

prior example of complete coupl<strong>in</strong>g, for nearest neighbor coupl<strong>in</strong>g the two<br />

3<br />

higher ly<strong>in</strong>g solutions are not degenerate. That is, the nearest neighbor<br />

coupl<strong>in</strong>g solutions differ from when all masses are l<strong>in</strong>early coupled.<br />

It is <strong>in</strong>terest<strong>in</strong>g to note that this example comb<strong>in</strong>es two coupl<strong>in</strong>g mechanisms<br />

that can be used to predict the solutions for two extreme cases<br />

by switch<strong>in</strong>g off one of these coupl<strong>in</strong>g mechanisms. Switch<strong>in</strong>g off the<br />

coupl<strong>in</strong>g spr<strong>in</strong>gs, by sett<strong>in</strong>g =0, makes all three normal frequencies<br />

degenerate with 1 = 2 = 3 = p <br />

<br />

. This corresponds to three <strong>in</strong>dependent<br />

identical pendula each with frequency = p <br />

Normal modes of three plane<br />

<br />

. Also the three pendula with nearest-neighbour<br />

l<strong>in</strong>ear comb<strong>in</strong>ations 1 2 3 also have this same frequency, <strong>in</strong> particular<br />

coupl<strong>in</strong>g.<br />

1 corresponds to an <strong>in</strong>-phase oscillation of the three pendula. The three<br />

uncoupled pendula are <strong>in</strong>dependent and any comb<strong>in</strong>ation the three modes is allowed s<strong>in</strong>ce the three frequencies<br />

are degenerate.<br />

The other extreme is to let <br />

=0 that is switch off the gravitational field or let →∞, then the only<br />

coupl<strong>in</strong>g is due to the two spr<strong>in</strong>gs. This results <strong>in</strong> 1 =0because there is no restor<strong>in</strong>g force act<strong>in</strong>g on the<br />

coherent motion of the three <strong>in</strong>-phase coupled oscillators; as a result, oscillatory motion cannot be susta<strong>in</strong>ed<br />

s<strong>in</strong>ce it corresponds to the center of mass oscillation with no external forces act<strong>in</strong>g which is spurious. That<br />

is, this spurious solution corresponds to constant l<strong>in</strong>ear translation.


378 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

14.8 Example: System of three bodies coupled by six spr<strong>in</strong>gs<br />

Consider the completely-coupled mechanical system shown <strong>in</strong> the adjacent figure.<br />

1) The first stage is to determ<strong>in</strong>e the potential and k<strong>in</strong>etic energies us<strong>in</strong>g an appropriate set of<br />

generalized coord<strong>in</strong>ates, which here are 1 and 2 . The potential energy is the sum of the potential energies<br />

for each of the six spr<strong>in</strong>gs<br />

whilethek<strong>in</strong>eticenergyisgivenby<br />

= 3 2 2 1 + 3 2 2 2 + 3 2 2 3 − 1 2 − 1 3 − 2 3<br />

= 1 2 ˙2 1 + 1 2 ˙2 2 + 1 2 ˙2 3<br />

2) Thesecondstageistoevaluatethepotentialenergy and<br />

k<strong>in</strong>etic energy tensors.<br />

⎧<br />

⎫<br />

⎧<br />

⎫<br />

⎨ 3 − − ⎬<br />

⎨ 0 0 ⎬<br />

V = − 3 −<br />

T = 0 0<br />

⎩<br />

⎭<br />

⎩<br />

⎭<br />

− − 3<br />

0 0 <br />

Note that for this case the k<strong>in</strong>etic energy tensor is diagonal whereas<br />

the potential energy tensor is nondiagonal and corresponds to complete<br />

coupl<strong>in</strong>g of the three coord<strong>in</strong>ates.<br />

3) The third stage is to use the potential and k<strong>in</strong>etic <br />

energy tensors to evaluate the secular determ<strong>in</strong>ant giv<strong>in</strong>g<br />

¡ ¢ k<br />

3 − <br />

2<br />

¡ − −<br />

−<br />

¢ 3 − <br />

2<br />

¡ −<br />

¯ − −<br />

¢ =0<br />

3 − <br />

2<br />

¯¯¯¯¯¯ System of three bodies coupled by six<br />

The expansion of this secular determ<strong>in</strong>ant yields<br />

spr<strong>in</strong>gs.<br />

¡<br />

− <br />

2 ¢¡ 4 − 2¢¡ 4 − 2¢ =0<br />

k<br />

k<br />

k<br />

k<br />

m<br />

m<br />

m<br />

k<br />

The solution for this complete-coupled system has two degenerate eigenvalues.<br />

r <br />

1 = 2 =2<br />

<br />

r <br />

3 =<br />

<br />

4) The fourth step is to <strong>in</strong>sert these eigenfrequencies <strong>in</strong>to the secular equation<br />

X ¡<br />

− 2 ¢<br />

=0<br />

<br />

to determ<strong>in</strong>e the coefficients <br />

5) The f<strong>in</strong>al stage is to write the general coord<strong>in</strong>ates <strong>in</strong> terms of the normal coord<strong>in</strong>ates<br />

The result is that the angular frequency 3 = p <br />

<br />

corresponds to a normal mode for which the three<br />

masses oscillate <strong>in</strong> phase correspond<strong>in</strong>g to a center-of-mass oscillation with no relative motion of the masses.<br />

3 = 1 √<br />

3<br />

( 1 + 2 + 3 )<br />

For this coherent motion only one spr<strong>in</strong>g per mass is stretched result<strong>in</strong>g <strong>in</strong> the same frequency as one<br />

mass on a spr<strong>in</strong>g. The other two solutions correspond to the three masses oscillat<strong>in</strong>g out of phase which<br />

implies all three spr<strong>in</strong>gs are stretched and thus the angular frequency is higher. S<strong>in</strong>ce the two eigenvalues<br />

1 = 2 =2 p <br />

<br />

are degenerate then there are only five <strong>in</strong>dependent equations to specify the six unknowns<br />

for the degenerate eigenvalues. Thus it is possible to select a comb<strong>in</strong>ation of the eigenvectors 1 and 2 such<br />

that the comb<strong>in</strong>ation is orthogonal to 3 Choose 31 =0to removes the <strong>in</strong>determ<strong>in</strong>acy. Then add<strong>in</strong>g or<br />

subtract<strong>in</strong>g gives that the normal modes are<br />

1 = √ 1 ( 1 − 2 +0) 2 = √ 1 ( 1 + 2 − 2 3 )<br />

2 2<br />

These two degenerate normal modes correspond to relative motion of the masses with stationary center-ofmass.


14.9. MOLECULAR COUPLED OSCILLATOR SYSTEMS 379<br />

14.9 Molecular coupled oscillator systems<br />

There are many examples of coupled oscillations <strong>in</strong> atomic and molecular physics most of which <strong>in</strong>volve<br />

nearest-neighbor coupl<strong>in</strong>g. The follow<strong>in</strong>g two examples are for molecular coupled oscillators. The triatomic<br />

molecule is a typical l<strong>in</strong>early-coupled molecular oscillator. The benzene molecule is an elementary example<br />

of a r<strong>in</strong>g structure coupled oscillator.<br />

14.9 Example: L<strong>in</strong>ear triatomic molecular CO 2<br />

Molecules provide excellent examples of vibrational modes <strong>in</strong>volv<strong>in</strong>g nearest neighbor coupl<strong>in</strong>g. Depend<strong>in</strong>g<br />

on the atomic structure, triatomic molecules can be either l<strong>in</strong>ear, like 2 ,orbentlikewater, 2 which<br />

has a bend angle of = 109 ◦ Amoleculewith atoms has 3 degrees of freedom. There are three degrees<br />

of freedom for translation and three degrees of freedom for rotation leav<strong>in</strong>g 3 − 6 degrees of freedom for<br />

vibrations. A triatomic molecule has three vibrational modes, two longitud<strong>in</strong>al and one transverse. Consider<br />

the normal modes for vibration of the l<strong>in</strong>ear molecule 2<br />

Longitud<strong>in</strong>al modes<br />

The coord<strong>in</strong>ate system used is illustrated <strong>in</strong> the adjacent figure.<br />

The Lagrangian for this system is<br />

µ <br />

=<br />

2 ˙2 1 + 2 ˙2 2 + <br />

2 ˙2 3 − 2 [( 2 − 1 ) 2 +( 3 − 2 ) 2 ]<br />

Evaluat<strong>in</strong>g the k<strong>in</strong>etic energy tensor gives<br />

⎧<br />

⎨<br />

T =<br />

⎩<br />

0 0<br />

0 0<br />

0 0 <br />

⎫<br />

⎬<br />

⎭<br />

while the potential energy tensor gives<br />

⎧<br />

⎨<br />

V = <br />

⎩<br />

1 −1 0<br />

−1 2 −1<br />

0 −1 1<br />

⎫<br />

⎬<br />

⎭<br />

The secular equation becomes<br />

¡ − 2 + ¢ ¡<br />

− 0<br />

− − 2 +2 ¢ ¡<br />

−<br />

¯ 0 − − 2 + <br />

¯¯¯¯¯¯ ¢ =0<br />

Note that the same answer is obta<strong>in</strong>ed us<strong>in</strong>g Newtonian mechanics. That is, the force equation gives<br />

¨ 1 − ( 2 − 1 ) = 0<br />

¨ 2 + ( 2 − 1 ) − ( 3 − 2 ) = 0<br />

¨ 3 − ( 3 − 2 ) = 0<br />

Let the solution be of the form<br />

Substitute this solution gives<br />

= =1 2 3<br />

¡<br />

− 2 + ¢ 1 − 2 = 0<br />

− 1 + ¡ − 2 +2 ¢ 2 − 3 = 0<br />

− 2 + ¡ 2 + ¢ 3 = 0<br />

This leads to the same secular determ<strong>in</strong>ant as given above with the matrix elements clustered along the<br />

diagonal for nearest-neighbor problems.


380 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

Expand<strong>in</strong>g the determ<strong>in</strong>ant and collect<strong>in</strong>g terms<br />

yields<br />

2 ¡ − 2 + ¢¡ − 2 + +2 ¢ =0<br />

Equat<strong>in</strong>g either of the three factors to zero gives<br />

1 = 0<br />

m K M K m<br />

x<br />

2 =<br />

3 =<br />

r <br />

<br />

s µ <br />

+ 2 <br />

<br />

The solutions are:<br />

1) 1 =0;This solution gives 1 = {1 1 1}. This<br />

mode is not an oscillation at all, but is a pure translation<br />

of the system as a whole as shown <strong>in</strong> the adjacent<br />

figure. There is no change <strong>in</strong> the restor<strong>in</strong>g forces s<strong>in</strong>ce<br />

the system moves such as not to change the length of the<br />

spr<strong>in</strong>gs, that is, they stay <strong>in</strong> their equilibrium positions.<br />

1<br />

2<br />

3<br />

4<br />

Normal modes of a l<strong>in</strong>ear triatomic molecule<br />

This motion corresponds to a spurious oscillation of the center of mass that results from referenc<strong>in</strong>g the<br />

three atom locations with respect to some fixed reference po<strong>in</strong>t. This reference po<strong>in</strong>t should have been chosen<br />

as the center of mass s<strong>in</strong>ce the motion of the center-of-mass already has been taken <strong>in</strong>to account separately.<br />

Spurious center of mass oscillations occur any time that the reference po<strong>in</strong>t is not at the center of mass for<br />

an isolated system with no external forces act<strong>in</strong>g.<br />

2) 2 = p <br />

: This solution corresponds to 2 = {1 0 −1} and is shown <strong>in</strong> the adjacent figure. The<br />

central mass rema<strong>in</strong>s stationary while the two end masses vibrate longitud<strong>in</strong>ally <strong>in</strong> opposite directions<br />

with the same amplitude. This mode has a stationary center of mass. For 2 the electrical geometry is<br />

− ++ − Mode 2 for 2 does not radiate electromagnetically because the center of charge is stationary<br />

with respect to the center of mass, that is, the electric dipole moment is constant.<br />

3) 3 =<br />

q ¡ <br />

+ 2<br />

<br />

¢<br />

: This solution corresponds to 3 = © 1 −2 ¡ <br />

<br />

¢<br />

1<br />

ª<br />

As shown <strong>in</strong> the adjacent<br />

figure, this motion corresponds to the two end masses vibrat<strong>in</strong>g <strong>in</strong> unison while the central mass vibrates<br />

oppositely with a different amplitude such that the center-of-mass is stationary. This 2 mode does radiate<br />

electromagnetically s<strong>in</strong>ce it corresponds to an oscillat<strong>in</strong>g electric dipole.<br />

It is <strong>in</strong>terest<strong>in</strong>g to note that the ratio 3<br />

2<br />

=1915 for 2 and the ratio of the two modes is <strong>in</strong>dependent<br />

of the potential energy tensor That is<br />

Transverse modes<br />

r<br />

³<br />

3<br />

= 1+2 ´<br />

2 <br />

The solutions qare:<br />

4) 4 = 2 ¡ ¢<br />

2+ <br />

This is the only non-spurious transverse mode 4 which corresponds to the two<br />

outside masses vibrat<strong>in</strong>g <strong>in</strong> unison transverse to the symmetry axis while the central mass vibrates oppositely.<br />

This mode radiates electric dipole radiation s<strong>in</strong>ce the electric dipole is oscillat<strong>in</strong>g.<br />

5) 5 =0. This transverse solution 5 has all three nuclei vibrat<strong>in</strong>g <strong>in</strong> unison transverse to the symmetry<br />

axis and corresponds to a spurious center of mass oscillation.<br />

6) 6 =0 This transverse solution 6 corresponds to a stationary central mass with the two outside<br />

masses vibrat<strong>in</strong>g oppositely. This corresponds to a rotational oscillation of the molecule which is spurious<br />

s<strong>in</strong>ce there are no torques act<strong>in</strong>g on the molecule for a central force. Rotational motion usually is taken <strong>in</strong>to<br />

account separately.<br />

The normal modes for the bent triatomic molecule are similar except that the oscillator coupl<strong>in</strong>g strength<br />

is reduced by the factor cos where is the bend angle.


14.9. MOLECULAR COUPLED OSCILLATOR SYSTEMS 381<br />

14.10 Example: Benzene r<strong>in</strong>g<br />

The benzene r<strong>in</strong>g comprises six carbon atoms bound <strong>in</strong> a plane hexagonal r<strong>in</strong>g. A classical analog of the<br />

benzener<strong>in</strong>gcomprises6 identical masses on a frictionless r<strong>in</strong>g bound by 6 identical spr<strong>in</strong>gs with l<strong>in</strong>ear<br />

spr<strong>in</strong>g constant as illustrated <strong>in</strong> the adjacent figure Consider only the <strong>in</strong>-plane motion, then the k<strong>in</strong>etic<br />

energy is given by<br />

= 1 6X<br />

2 2<br />

The potential energy equals<br />

"<br />

= 1 6X<br />

6X<br />

#<br />

2 2 +1 − )<br />

=1( 2 = 2 2 − 1 2 − 2 3 − 3 4 − 4 5 − 5 6 − 6 1<br />

=1<br />

where =7≡ 1. Thus the k<strong>in</strong>etic energy and potential energy tensors are given by<br />

⎛<br />

⎞<br />

⎛<br />

⎞<br />

1 0 0 0 0 0<br />

2 −1 0 0 0 −1<br />

0 1 0 0 0 0<br />

−1 2 −1 0 0 0<br />

= 2 0 0 1 0 0 0<br />

⎜ 0 0 0 1 0 0<br />

= 2 0 −1 2 −1 0 0<br />

⎟<br />

⎜ 0 0 −1 2 −1 0<br />

⎟<br />

⎝ 0 0 0 0 1 0 ⎠<br />

⎝ 0 0 0 −1 2 −1 ⎠<br />

0 0 0 0 0 1<br />

−1 0 0 0 −1 2<br />

This nearest-neighbor system <strong>in</strong>cludes non-zero ( 1) and (1) elements due to the r<strong>in</strong>g structure. Def<strong>in</strong>e<br />

= 2<br />

<br />

− 2 then the solution of the set of l<strong>in</strong>ear homogeneous equations requires that<br />

1 0 0 0 1<br />

1 1 0 0 0<br />

0 1 1 0 0<br />

0 0 1 1 0<br />

=0<br />

0 0 0 1 1<br />

¯ 1 0 0 0 1 ¯<br />

that is<br />

=1<br />

( − 2) ( − 1) 2 ( +1) 2 ( +2)=0<br />

The eigenvalues and eigenfunctions are given <strong>in</strong> the table<br />

<strong>Classical</strong> analog of a benzene molecular r<strong>in</strong>g.<br />

n x 2 Normal modes<br />

1 2<br />

4<br />

<br />

1 − 2 + 3 − 4 + 5 − 6<br />

2 1<br />

3<br />

<br />

− 1 + 3 − 4 + 6<br />

3 1<br />

3<br />

<br />

− 1 + 2 − 4 + 5<br />

<br />

4 −1<br />

<br />

1 − 3 − 4 + 6<br />

<br />

5 −1<br />

<br />

− 1 − 2 + 4 + 5<br />

K 5 6 K<br />

6 −2 0 1 + 2 + 3 + 4 + 5 + 6<br />

m m<br />

K<br />

Note the follow<strong>in</strong>g properties of the normal modes and their frequencies.<br />

=1: Adjacent masses vibrate 180 ◦ out of phase, thus each spr<strong>in</strong>g has maximal compression or extension,<br />

lead<strong>in</strong>g to the energy of this normal mode be<strong>in</strong>g the highest.<br />

=2 3: These two solutions are degenerate and correspond to two pairs of masses vibrat<strong>in</strong>g out of phase<br />

while the third pair of masses are stationary. Thus the energy of this normal mode is slightly lower than the<br />

=1normal mode. Any comb<strong>in</strong>ation of these degenerate normal modes are equally good solutions.<br />

=4 5: From the figure it can be seen that both of these solutions correspond to a center of mass<br />

oscillation and thus these modes are spurious.<br />

=6: This vibrational mode has zero energy correspond<strong>in</strong>g to zero restor<strong>in</strong>g force and all six masses<br />

mov<strong>in</strong>g uniformly <strong>in</strong> the same direction. This mode corresponds to the rotation of the benzene molecule about<br />

the symmetry axis of the r<strong>in</strong>g which usually is taken <strong>in</strong>to account assum<strong>in</strong>g a separate rotational component.<br />

This classical analog of the benzene molecule is <strong>in</strong>terest<strong>in</strong>g because it simultaneously exhibits degenerate<br />

normal modes, spurious center of mass oscillation, and a rotational mode.<br />

˙ 2 <br />

m<br />

K<br />

m<br />

4<br />

3<br />

K<br />

2<br />

m<br />

1<br />

K<br />

m


382 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

14.10 Discrete Lattice Cha<strong>in</strong><br />

Crystall<strong>in</strong>e lattices and l<strong>in</strong>ear molecules are important classes of coupled oscillator systems where nearest<br />

neighbor <strong>in</strong>teractions dom<strong>in</strong>ate. A crystall<strong>in</strong>e lattice comprises thousands of coupled oscillators <strong>in</strong> a threedimensional<br />

matrix with atomic spac<strong>in</strong>g of a few 10 −10 . Even though a full description of the dynamics of<br />

crystall<strong>in</strong>e lattices demands a quantal treatment, a classical treatment is of <strong>in</strong>terest s<strong>in</strong>ce classical mechanics<br />

underlies many features of the motion of atoms <strong>in</strong> a crystall<strong>in</strong>e lattice. The l<strong>in</strong>ear discrete lattice cha<strong>in</strong> is<br />

the simplest example of many-body coupled oscillator systems that can illum<strong>in</strong>ate the physics underly<strong>in</strong>g a<br />

range of <strong>in</strong>terest<strong>in</strong>g phenomena <strong>in</strong> solid-state physics. As illustrated <strong>in</strong> example 27 the l<strong>in</strong>ear approximation<br />

usually is applicable for small-amplitude displacements of nearest-neighbor <strong>in</strong>teract<strong>in</strong>g systems which<br />

greatly simplifies treatment of the lattice cha<strong>in</strong>. The l<strong>in</strong>ear discrete lattice cha<strong>in</strong> <strong>in</strong>volves three <strong>in</strong>dependent<br />

polarization modes, one longitud<strong>in</strong>al mode, plus two perpendicular transverse modes. The 3 degrees of<br />

freedom for the atoms, on a discrete l<strong>in</strong>ear lattice cha<strong>in</strong>, are partitioned with degrees of freedom for each<br />

of the three polarization modes. These three polarization modes each have normal modes, or travell<strong>in</strong>g<br />

waves, and exhibit quantization, dispersion, and can have a complex wave number.<br />

14.10.1 Longitud<strong>in</strong>al motion<br />

The equations of motion for longitud<strong>in</strong>al modes of the lattice cha<strong>in</strong> can be derived by consider<strong>in</strong>g a l<strong>in</strong>ear<br />

cha<strong>in</strong> of identical masses, of mass separated by a uniform spac<strong>in</strong>g asshown<strong>in</strong>Fig147. Assume<br />

that the masses are coupled by +1 spr<strong>in</strong>gs, with spr<strong>in</strong>g constant , where both ends of the cha<strong>in</strong> are<br />

fixed, that is, the displacements 0 = +1 =0and velocities ˙ 0 = ˙ +1 =0 The force required to stretch a<br />

length of the cha<strong>in</strong> a longitud<strong>in</strong>al displacements, for mass is = Thus the potential energy for<br />

stretch<strong>in</strong>g the spr<strong>in</strong>g for segment ( −1 − ) is = 2 ( −1 − ). The total potential and k<strong>in</strong>etic energies<br />

are<br />

= +1<br />

X<br />

( −1 − ) 2 (14.74)<br />

2<br />

=1<br />

= 1 2 X<br />

˙ 2 (14.75)<br />

=1<br />

S<strong>in</strong>ce ˙ +1 =0the k<strong>in</strong>etic energy and Lagrangian can be<br />

extended to = +1, that is, the Lagrangian can be written<br />

as<br />

= 1 +1<br />

X ³ 2´<br />

˙ 2 − ( −1 − ) (14.76)<br />

2<br />

=1<br />

Us<strong>in</strong>g this Lagrangian <strong>in</strong> the Lagrange-Euler equations<br />

gives the follow<strong>in</strong>g second-order equation of motion for longitud<strong>in</strong>al<br />

oscillations<br />

¨ = 2 ( −1 − 2 + +1 ) (14.77)<br />

where =1 2 and where<br />

r <br />

≡<br />

<br />

14.10.2 Transverse motion<br />

(14.78)<br />

d d d d<br />

q j-2<br />

q j-1<br />

q j<br />

q j+1<br />

q j+2<br />

Figure 14.7: Portion of a lattice cha<strong>in</strong> of identical<br />

masses connected by identical spr<strong>in</strong>gs<br />

of spr<strong>in</strong>g constant . The displacement of the<br />

mass from the equilibrium position is <br />

assumed to be positive to the right.<br />

The equations of motion for transverse motion on a l<strong>in</strong>ear discrete lattice cha<strong>in</strong>, illustrated <strong>in</strong> figure 148,<br />

can be derived by consider<strong>in</strong>g the displacements of the mass for identical masses, with mass <br />

separated by equal spac<strong>in</strong>gs and assum<strong>in</strong>g that the tension <strong>in</strong> the str<strong>in</strong>g is = ¡ ¢<br />

<br />

. Assum<strong>in</strong>g that the<br />

transverse deflections are small, then the − 1 to spr<strong>in</strong>g is stretched to a length<br />

q<br />

0 = 2 +( − −1 ) 2 (14.79)


14.10. DISCRETE LATTICE CHAIN 383<br />

Thus the <strong>in</strong>cremental stretch<strong>in</strong>g is<br />

∼ ( − −1 ) 2<br />

2<br />

The work done aga<strong>in</strong>st the tension is · per segment. Thus the<br />

total potential energy is<br />

(14.80)<br />

= +1<br />

X<br />

( −1 − ) 2 (14.81)<br />

2<br />

=1<br />

where 0 and +1 are identically zero.<br />

The k<strong>in</strong>etic energy is<br />

= 1 <br />

2 X<br />

˙ 2 (14.82)<br />

S<strong>in</strong>ce ˙ +1 =0 the k<strong>in</strong>etic energy and Lagrangian summations can<br />

be extended to = +1,thatis<br />

= 1 +1<br />

X ³ ˙ 2 − 2´<br />

2<br />

( −1 − ) (14.83)<br />

=1<br />

=1<br />

d d d d<br />

Figure 14.8: Transverse motion of a<br />

l<strong>in</strong>ear discrete lattice cha<strong>in</strong><br />

Us<strong>in</strong>g this Lagrangian <strong>in</strong> the Lagrange Euler equations gives the follow<strong>in</strong>g second-order equation of motion<br />

for transverse oscillations<br />

¨ = 2 ( −1 − 2 + +1 ) (14.84)<br />

where =1 2 and<br />

r <br />

≡<br />

(14.85)<br />

<br />

The normal modes for the transverse modes comprise stand<strong>in</strong>g waves that satisfy the same boundary<br />

conditions as for the longitud<strong>in</strong>al modes. The equations of motion for longitud<strong>in</strong>al motion, equation<br />

1477 or transverse motion, equation 1484 are identical <strong>in</strong> form. The major difference is that 0 for the<br />

transverse normal modes ≡ p <br />

differs from that for the longitud<strong>in</strong>al modes which is ≡ p <br />

.Thus<br />

the follow<strong>in</strong>g discussion of the normal modes on a discrete lattice cha<strong>in</strong> is identical <strong>in</strong> form for both transverse<br />

and longitud<strong>in</strong>al waves.<br />

14.10.3 Normal modes<br />

The normal modes of the equations of motion on the discrete lattice cha<strong>in</strong>, are either longitud<strong>in</strong>al or<br />

transverse stand<strong>in</strong>g waves that satisfy the boundary conditions at the extreme ends of the lattice cha<strong>in</strong>.<br />

The solutions can be given by assum<strong>in</strong>g that the identical masses on the cha<strong>in</strong> oscillate with a common<br />

frequency . Then the displacement amplitude for the mass can be written <strong>in</strong> the form<br />

() = (14.86)<br />

where the amplitude can be complex. Substitution <strong>in</strong>to the preced<strong>in</strong>g equations of motion, 1477 1484<br />

yields the follow<strong>in</strong>g recursion relation<br />

¡<br />

− 2 +2 2 ¢<br />

− 2 0 ( −1 + +1 )=0 (14.87)<br />

where =1 2 Note that the boundary conditions, 0 =0and +1 =0require that = +1 =0<br />

The above recursion relation corresponds to a system of homogeneous algebraic equations with <br />

unknowns 1 2 A non-trivial solution is given by sett<strong>in</strong>g the determ<strong>in</strong>ant of its coefficients equal to<br />

zero<br />

−<br />

¯¯¯¯¯¯¯¯¯¯<br />

2 +2 2 − 2 0 0<br />

− 2 − 2 +2 2 − 2 0<br />

0 − 2 − 2 +2 2 − 2 <br />

=0 (14.88)<br />

<br />

0 0 − 2 − 2 +2 2 ¯


384 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

This secular determ<strong>in</strong>ant corresponds to the special case of nearest neighbor <strong>in</strong>teractions with the k<strong>in</strong>etic<br />

energy tensor T be<strong>in</strong>g diagonal and the potential energy tensor V <strong>in</strong>volv<strong>in</strong>g coupl<strong>in</strong>g only to adjacent<br />

masses. The secular determ<strong>in</strong>ant is of order and thus determ<strong>in</strong>es exactly eigen frequencies for each<br />

polarization mode.<br />

For large the solution of this problem is more efficiently obta<strong>in</strong>ed by us<strong>in</strong>g a recursion relation approach,<br />

rather than solv<strong>in</strong>g the above secular determ<strong>in</strong>ant. The trick is to assume that the phase differences <br />

between the motion of adjacent masses all are identical for a given polarization. Then the amplitude for the<br />

mass for the frequency mode is of the form<br />

= ( − )<br />

(14.89)<br />

Insert the above <strong>in</strong>to the recursion relation (1487) gives<br />

¡<br />

−<br />

2<br />

+2 2 ¢ £ ¤<br />

− <br />

2<br />

0 <br />

− + <br />

=0 (14.90)<br />

which reduces to<br />

2 =2 2 − 2 2 cos =4 2 s<strong>in</strong> 2 <br />

2<br />

that is<br />

=2 s<strong>in</strong> <br />

(14.91)<br />

2<br />

where =1 2 3 <br />

Now it is necessary to determ<strong>in</strong>e the phase angle which can be done by apply<strong>in</strong>g the boundary<br />

conditions for stand<strong>in</strong>g waves on the lattice cha<strong>in</strong>. These boundary conditions for stationary modes require<br />

that the ends of the lattice cha<strong>in</strong> are nodes, that is = (+1) =0 Us<strong>in</strong>g the fact that only the real<br />

part of has physical mean<strong>in</strong>g, leads to the amplitude for the mass for the mode to be<br />

= cos ( − ) (14.92)<br />

The boundary condition 0 =0requires that the phase = 2<br />

That is<br />

³<br />

= cos − ´<br />

= s<strong>in</strong> <br />

2<br />

(14.93)<br />

where =1 2 <br />

The boundary condition for = +1 gives<br />

(+1) =0= s<strong>in</strong> ( +1) (14.94)<br />

Therefore<br />

where =1 2 3 . Thatis<br />

( +1) = (14.95)<br />

=<br />

<br />

+1 = <br />

( +1) = <br />

<br />

= <br />

2<br />

(14.96)<br />

where =( +1) is the total length of the discrete lattice cha<strong>in</strong>.<br />

The eigen frequencies for a given polarization are given by<br />

<br />

=2 s<strong>in</strong><br />

2( +1) =2 <br />

s<strong>in</strong><br />

2( +1) =2 s<strong>in</strong> <br />

2 =2 s<strong>in</strong> <br />

2<br />

where the correspond<strong>in</strong>g wavenumber is given by<br />

(14.97)<br />

=<br />

<br />

( +1) = = 2<br />

<br />

(14.98)<br />

This implies that the normal modes are quantized with half-wavelengths <br />

2<br />

= .


14.10. DISCRETE LATTICE CHAIN 385<br />

r = 1 r = 2<br />

1.0<br />

1.0<br />

0.8<br />

0.8<br />

0.6<br />

0.6<br />

0.4<br />

0.4<br />

0.2<br />

0.2<br />

0.0<br />

-0.2<br />

0. 1 0.2 0. 3 0. 4 0. 5 0. 6 0. 7 0. 8 0. 9 1. 0<br />

0.0<br />

-0.2<br />

0. 1 0.2 0. 3 0. 4 0. 5 0. 6 0. 7 0. 8 0. 9 1. 0<br />

-0.4<br />

-0.4<br />

-0.6<br />

-0.6<br />

-0.8<br />

-0.8<br />

-1.0<br />

-1.0<br />

r = 3 r = 4<br />

1.0<br />

1.0<br />

0.8<br />

0.8<br />

0.6<br />

0.6<br />

0.4<br />

0.4<br />

0.2<br />

0.2<br />

0.0<br />

-0.2<br />

0. 1 0.2 0. 3 0. 4 0. 5 0. 6 0. 7 0. 8 0. 9 1. 0<br />

0.0<br />

-0.2<br />

0. 1 0.2 0. 3 0. 4 0. 5 0. 6 0. 7 0. 8 0. 9 1. 0<br />

-0.4<br />

-0.4<br />

-0.6<br />

-0.6<br />

-0.8<br />

-0.8<br />

-1.0<br />

-1.0<br />

r = 5 r = 6<br />

1.0<br />

1.0<br />

0.8<br />

0.8<br />

0.6<br />

0.6<br />

0.4<br />

0.4<br />

0.2<br />

0.2<br />

0.0<br />

-0.2<br />

0. 1 0.2 0. 3 0. 4 0. 5 0. 6 0. 7 0. 8 0. 9 1. 0<br />

0.0<br />

-0.2<br />

0. 1 0.2 0. 3 0. 4 0. 5 0. 6 0. 7 0. 8 0. 9 1. 0<br />

-0.4<br />

-0.4<br />

-0.6<br />

-0.6<br />

-0.8<br />

-0.8<br />

-1.0<br />

-1.0<br />

Figure 14.9: Plots of the maximal vibrational amplitudes for the frequency s<strong>in</strong>usoidal mode, versus<br />

distance along the cha<strong>in</strong>, for transverse normal modes of a vibrat<strong>in</strong>g discrete lattice with =5. Only =<br />

1 2 3 4 5 are dist<strong>in</strong>ct modes because =6is a null mode. Note that the modes with =7 8 9 10 11 12<br />

shown dashed, duplicate the locations of the mass displacement given by the lower-order modes.<br />

Comb<strong>in</strong><strong>in</strong>g equations 1496 and 1493 gives the maximum amplitudes for the eigenvectors to be<br />

= s<strong>in</strong> <br />

(14.99)<br />

2<br />

For <strong>in</strong>dependent l<strong>in</strong>ear oscillators there are only <strong>in</strong>dependent normal modes, that is, for = +1the<br />

s<strong>in</strong>e function <strong>in</strong> equation 1497 must be zero. Beyond = the equations do not describe physically new<br />

situations. This is illustrated by figure 149 which shows the transverse modes of a lattice cha<strong>in</strong> with =5.<br />

There are only =5<strong>in</strong>dependent normal modes of this system s<strong>in</strong>ce = +1=6corresponds to a null<br />

mode with all () =0. Also note that the solutions for +1 shown dashed, replicate the mass<br />

locations of modes with +1, that is, the modes with 6 are replicas of the lower-order modes.<br />

Note that has a maximum value ≤ 2 0 s<strong>in</strong>ce the s<strong>in</strong>e function cannot exceed unity. This leads<br />

to a maximum frequency =2 0 called the cut-off frequency, which occurs when = . That is, the<br />

null-mode occurs when = +1 for which equation 1499 equals zero The range of quantized normal<br />

modes that can occur is <strong>in</strong>tuitive. That is, the longest half-wavelength max<br />

2<br />

= =( +1) equals the total<br />

length of the discrete lattice cha<strong>in</strong>. The shortest half-wavelength −<br />

2<br />

= is set by the lattice spac<strong>in</strong>g.<br />

Thus the discrete wavenumbers of the normal modes, for each polarization, range from 1 to 1 where is<br />

an <strong>in</strong>teger.<br />

Assum<strong>in</strong>g real the normal coord<strong>in</strong>ate and correspond<strong>in</strong>g frequency are,<br />

= <br />

(14.100)<br />

Equations 1497 and 1499 give the angular frequency and displacement. Note that superposition applies<br />

s<strong>in</strong>ce this system is l<strong>in</strong>ear. Therefore the most general solution for each polarization can be any superposition<br />

of the form<br />

() =<br />

X<br />

=1<br />

∙<br />

s<strong>in</strong><br />

<br />

( +1)<br />

¸<br />

(14.101)


386 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

14.10.4 Travell<strong>in</strong>g waves<br />

Travell<strong>in</strong>g waves are equally good solutions of the equations of motion 1477 1484 as are the normal modes.<br />

Travell<strong>in</strong>g waves on the one-dimensional lattice cha<strong>in</strong> will be of the form<br />

( ) = (±) (14.102)<br />

where the distance along the cha<strong>in</strong> = , that is, it is quantized <strong>in</strong> units of the cell spac<strong>in</strong>g , with be<strong>in</strong>g<br />

an <strong>in</strong>teger. The positive sign <strong>in</strong> the exponent corresponds to a wave travell<strong>in</strong>g <strong>in</strong> the − direction while<br />

the negative sign corresponds to a wave travell<strong>in</strong>g <strong>in</strong> the + direction. The velocity of a fixed phase of the<br />

travell<strong>in</strong>g wave must satisfy that ± is a constant. This will occur if the phase velocity of the wave is<br />

given by<br />

= <br />

= (14.103)<br />

<br />

Thewavehasafrequency = 2 and wavelength = 2 thus the phase velocity = = .<br />

Insert<strong>in</strong>g the travell<strong>in</strong>g wave 14102 <strong>in</strong>to the transverse equation of motion 1484 for the discrete lattice<br />

cha<strong>in</strong> gives<br />

− 2 = 2 0( − − 2+<br />

) (14.104)<br />

where =1 2 That is<br />

= ±2 0 s<strong>in</strong> <br />

(14.105)<br />

2<br />

The phase is determ<strong>in</strong>ed by the Born-von Karman periodic boundary condition that assumes that the<br />

cha<strong>in</strong> is duplicated <strong>in</strong>def<strong>in</strong>itely on either side of = ± <br />

. Thus, for discrete masses, must satisfy the<br />

condition that = + .Thatis<br />

=1 (14.106)<br />

That is<br />

= 2<br />

(14.107)<br />

<br />

Note that the periodic boundary condition gives discrete modes<br />

for wavenumbers between<br />

where the <strong>in</strong>dex<br />

− ≤ ≤ + <br />

(14.108)<br />

First Brillou<strong>in</strong> zone<br />

Thus equation 14105 becomes<br />

= − 2 − 2 +1 2 − 1 2<br />

= ±2 0 s<strong>in</strong> <br />

(14.109)<br />

2<br />

Equation 14109 is a dispersion relation that is identical to equation<br />

1497 derived dur<strong>in</strong>g the discussion of the normal modes of the<br />

lattice cha<strong>in</strong>. This confirms that the travell<strong>in</strong>g waves on the lattice<br />

cha<strong>in</strong> are equally good solutions as the normal stand<strong>in</strong>g-wave<br />

modes. Clearly, superposition of the stand<strong>in</strong>g-wave normal modes<br />

can lead to travell<strong>in</strong>g waves and vice versa.<br />

14.10.5 Dispersion<br />

0<br />

Figure 14.10: Plot of the dispersion<br />

curve ( versus ) for a monoatomic<br />

l<strong>in</strong>ear lattice cha<strong>in</strong> subject to only<br />

nearest neighbor <strong>in</strong>teractions. The<br />

first Brillou<strong>in</strong> zone is the segment between<br />

− ≤ ≤ <br />

which covers all<br />

<strong>in</strong>dependent solutions.<br />

The lattice cha<strong>in</strong> is an <strong>in</strong>terest<strong>in</strong>g example of a dispersive system <strong>in</strong> that is a function of Figure 1410<br />

shows a plot of the dispersion curve ( versus ) for a monoatomic l<strong>in</strong>ear lattice cha<strong>in</strong> subject to only nearest<br />

neighbor <strong>in</strong>teractions. Note that depends l<strong>in</strong>early on for small and that <br />

<br />

=0at the boundaries of<br />

the first Brillou<strong>in</strong> zone.<br />

kd


14.10. DISCRETE LATTICE CHAIN 387<br />

The lattice cha<strong>in</strong> has a phase velocity for the wave given by<br />

<br />

= ¯<br />

<br />

<br />

¯s<strong>in</strong> ¯<br />

2<br />

= 0 <br />

<br />

while the group velocity is<br />

<br />

=<br />

<br />

2<br />

µ <br />

= 0 cos <br />

<br />

<br />

2<br />

(14.110)<br />

(14.111)<br />

Note that <strong>in</strong> the limit when <br />

2<br />

→ 0 the phase velocity and group velocity are identical, that is, <br />

= 0 <br />

<br />

<br />

14.10.6 Complex wavenumber<br />

=<br />

The maximum allowed frequency, which is called the cut-off frequency, =2 0 occurs when = , that<br />

is, 2 = . That is, the m<strong>in</strong>imum half-wavelength equals the spac<strong>in</strong>g between the discrete masses. At the<br />

cut-off frequency, the phase velocity is <br />

= 2 0 and the group velocity =0<br />

It is <strong>in</strong>terest<strong>in</strong>g to note that can exceed the cut-off frequency =2 0 if is assumed to be complex,<br />

that is, if<br />

= − Γ (14.112)<br />

Then<br />

=2 0 s<strong>in</strong> <br />

2 =2 0 s<strong>in</strong> µ<br />

2 ( − Γ )=2 0 s<strong>in</strong> <br />

2 cosh Γ <br />

2 − cos <br />

2 s<strong>in</strong>h Γ <br />

<br />

(14.113)<br />

2<br />

To ensure that is real, the imag<strong>in</strong>ary term must be zero, that is<br />

cos <br />

=0 (14.114)<br />

2<br />

Therefore<br />

s<strong>in</strong> <br />

=1 (14.115)<br />

2<br />

that is, = , and the dispersion relation between and for 2 0 becomes<br />

=2 0 cosh Γ <br />

2<br />

which <strong>in</strong>creases with Γ. Thus,when =2 0 then the amplitude of the wave is of the form<br />

() = −Γ ( − )<br />

(14.116)<br />

(14.117)<br />

which corresponds to a spatially damped oscillatory wave with phase velocity<br />

<br />

<br />

= <br />

<br />

(14.118)<br />

and damp<strong>in</strong>g factor Γ .<br />

There are many examples <strong>in</strong> physics where the wavenumber is complex as exhibited by the discrete lattice<br />

cha<strong>in</strong> for 2<br />

≤ . Other examples are electromagnetic waves <strong>in</strong> conductors or plasma (example 35), matter<br />

waves tunnell<strong>in</strong>g through a potential barrier, or stand<strong>in</strong>g waves on musical <strong>in</strong>struments which have a complex<br />

wavenumber due to damp<strong>in</strong>g.<br />

This simple toy model of the discrete l<strong>in</strong>ear lattice cha<strong>in</strong> has illustrated that classical mechanics expla<strong>in</strong>s<br />

many features of the many-body nearest-neighbor coupled l<strong>in</strong>ear oscillator system, <strong>in</strong>clud<strong>in</strong>g normal modes,<br />

stand<strong>in</strong>g and travell<strong>in</strong>g waves, cut-off frequency dispersion, and complex wavenumber. These phenomena<br />

feature prom<strong>in</strong>ently <strong>in</strong> applications of the quantal discrete coupled-oscillator system to solid-state physics.


388 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

14.11 Damped coupled l<strong>in</strong>ear oscillators<br />

The discussion of coupled l<strong>in</strong>ear oscillators has neglected non-conservative damp<strong>in</strong>g forces which always exist<br />

to some extent <strong>in</strong> physical systems. In general, dissipative forces are non l<strong>in</strong>ear which greatly complicates<br />

solv<strong>in</strong>g the equations of motion for such coupled oscillator systems. However, for some systems the dissipative<br />

forces depend l<strong>in</strong>early on velocity which allows use of the Rayleigh dissipation function, described <strong>in</strong> chapter<br />

104. The most general def<strong>in</strong>ition of the Rayleigh dissipation function, 104 was given to be<br />

R = 1 2<br />

X<br />

=1 =1<br />

X<br />

˙ ˙ (14.119)<br />

For this special case, it was shown <strong>in</strong> chapter 10 that the Lagrange equations can be written <strong>in</strong> terms of the<br />

Rayleigh dissipation function as ½ µ <br />

− ¾<br />

+ R = (14.120)<br />

˙ ˙ <br />

where are generalized forces act<strong>in</strong>g on the system that are not absorbed <strong>in</strong>to the potential Us<strong>in</strong>g<br />

equations 1443 1444 and 14120 allows the equations of motion for damped coupled l<strong>in</strong>ear oscillators to<br />

bewritten<strong>in</strong>amatrixformas<br />

{T} ¨q + {C} ˙q+ {V} q = {Q} (14.121)<br />

where the symmetric matrices {T} {C} and {V} are positive def<strong>in</strong>ite for positive def<strong>in</strong>ite systems. Rayleigh<br />

po<strong>in</strong>ted out that <strong>in</strong> the special case where the damp<strong>in</strong>g matrix {C} is a l<strong>in</strong>ear comb<strong>in</strong>ation of the {T} and<br />

{V} matrices, then the matrix {C} is diagonal lead<strong>in</strong>g to a separation of the damped system <strong>in</strong>to normal<br />

modes. As discussed <strong>in</strong> chapter 4 many systems <strong>in</strong> nature are l<strong>in</strong>ear for small amplitude oscillations allow<strong>in</strong>g<br />

use of the Rayleigh dissipation function which provides an analytic solution. However, <strong>in</strong> general, except for<br />

when {C} is small, this separation <strong>in</strong>to normal modes is not possible for damped systems and the solutions<br />

must be obta<strong>in</strong>ed numerically.<br />

The follow<strong>in</strong>g example illustrates approaches used to handle l<strong>in</strong>early-damped coupled-oscillator systems.<br />

14.11 Example: Two l<strong>in</strong>early-damped coupled l<strong>in</strong>ear oscillators<br />

Consider the two coupled oscillator system shown<br />

where the two carts have spr<strong>in</strong>g constants 1 2 and<br />

l<strong>in</strong>ear damp<strong>in</strong>g constants 1 2 . As discussed <strong>in</strong> example<br />

143, the k<strong>in</strong>etic energy tensor is given by<br />

= 1 2 1 ˙ 1 2 + 1 2 2 ˙ 2<br />

2<br />

and the potential energy is given by<br />

= 1 h 1 1 2 + 2 ( 2 − 1 ) 2i<br />

2<br />

= 1 £<br />

(1 + 2 ) 1 2 − 2 2 1 2 + 2 2<br />

2<br />

2¤<br />

()<br />

()<br />

Two l<strong>in</strong>early-damped coupled l<strong>in</strong>ear oscillators.<br />

Similarly the Rayleigh dissipation function has the form<br />

R = 1 £<br />

1 ˙ 2 ¡<br />

1 + 2 ˙<br />

2<br />

2<br />

2 − ˙ 1¢¤ 2 1 £<br />

= (1 + 2 ) ˙ 1 2 − 2 2 ˙ 1 ˙ 2 + 2 ˙ 2<br />

2<br />

2¤<br />

Insert<strong>in</strong>g and <strong>in</strong>to equation 14120 gives the two equations of motion to be<br />

()<br />

1¨ 1 +( 1 + 2 ) ˙ 1 − 2 ˙ 2 +( 1 + 2 ) 1 − 2 2 = 0<br />

2¨ 2 − 2 ˙ 1 + 2 ˙ 2 − 2 1 + 2 2 = 0<br />

When the drag is zero the solution of these two coupled equations can be separated <strong>in</strong>to two <strong>in</strong>dependent<br />

normal modes of the system as described earlier. Usually it is not possible to separate the motion <strong>in</strong>to<br />

decoupled normal modes except for certa<strong>in</strong> cases where the dissipative forces can be described by Rayleigh’s<br />

dissipation function.


14.12. COLLECTIVE SYNCHRONIZATION OF COUPLED OSCILLATORS 389<br />

14.12 Collective synchronization of coupled oscillators<br />

Collective synchronization of coupled oscillators is a multifaceted phenomenon where large ensembles of<br />

coupled oscillators, with comparable natural frequencies, self synchronize lead<strong>in</strong>g to coherent collective modes<br />

of motion. Biological examples <strong>in</strong>clude congregations of synchronously flash<strong>in</strong>g fireflies, crickets that chirp <strong>in</strong><br />

unison, an audience clapp<strong>in</strong>g at the end of a performance, networks of pacemaker cells <strong>in</strong> the heart, <strong>in</strong>sul<strong>in</strong>secret<strong>in</strong>g<br />

cells <strong>in</strong> the pancreas, as well as neural networks <strong>in</strong> the bra<strong>in</strong> and sp<strong>in</strong>al cord that control rhythmic<br />

behaviors such as breath<strong>in</strong>g, walk<strong>in</strong>g, and eat<strong>in</strong>g. Example 1413 illustrates an application to nuclei.<br />

An ensemble of coupled oscillators will have a frequency distribution with a f<strong>in</strong>ite width. It is <strong>in</strong>terest<strong>in</strong>g<br />

to elucidate how an ensemble of coupled oscillators, that have a f<strong>in</strong>ite width frequency distribution, can self<br />

synchronize their motion to a unique common frequency, and how that synchronization is ma<strong>in</strong>ta<strong>in</strong>ed over<br />

long time periods. The answers to these issues provide <strong>in</strong>sight <strong>in</strong>to the dynamics of coupled oscillators.<br />

The discussion of coupled oscillators has implicitly assumed identical undamped l<strong>in</strong>ear oscillators that<br />

have identical, <strong>in</strong>f<strong>in</strong>itely-sharp, natural frequencies . In nature typical coupled oscillators can have a f<strong>in</strong>itewidth<br />

frequency distribution () about some average value, due to the natural variability of the oscillator<br />

parameters for biological systems, the manufactur<strong>in</strong>g tolerances for mechanical oscillators, or the natural<br />

Lorentzian frequency distribution associated with the uncerta<strong>in</strong>ty pr<strong>in</strong>ciple that occurs even for atomic clocks<br />

where the oscillator frequencies are def<strong>in</strong>ed directly by the physical constants. Assume that the ensemble of<br />

coupled oscillators has a frequency distribution () about some average value.<br />

Undamped l<strong>in</strong>ear oscillators have elliptical closed-path trajectories <strong>in</strong> phase space whereas dissipation<br />

leads to a spiral attractor unless the system is driven such as to preserve the total energy. As described<br />

<strong>in</strong> chapter 44 many systems <strong>in</strong> nature, especially biological systems, have closed limit cycles <strong>in</strong> phase<br />

space where the energy lost to dissipation is replenished by a driv<strong>in</strong>g mechanism. The simplest systems for<br />

understand<strong>in</strong>g collective synchronization of coupled oscillators are those that <strong>in</strong>volve closed limit cycles <strong>in</strong><br />

phase space.<br />

N. Wiener first recognized the ubiquity of collective synchronization <strong>in</strong> the natural world, but his mathematical<br />

approach, based on Fourier <strong>in</strong>tegrals, was not suited to this problem. A more fruitful approach was<br />

pioneered <strong>in</strong> 1975 by an undergraduate student A.T. W<strong>in</strong>free[W<strong>in</strong>67] who recognized that the long-time behavior<br />

of a large ensemble of limit-cycle oscillators can be characterized <strong>in</strong> the simplest terms by consider<strong>in</strong>g<br />

only the phase of closed phase-space trajectories. He assumed that the <strong>in</strong>stantaneous state of an ensemble<br />

of oscillators can be represented by po<strong>in</strong>ts distributed around the circular phase-space diagram shown <strong>in</strong><br />

figure 1411. For uncoupled oscillators these po<strong>in</strong>ts will be distributed randomly around the circle, whereas<br />

coupl<strong>in</strong>g of the oscillators will result <strong>in</strong> a spatial correlation of the po<strong>in</strong>ts. That is, the dynamics of the<br />

phases can be visualized as a swarm of po<strong>in</strong>ts runn<strong>in</strong>g around the unit circle <strong>in</strong> the complex plane of the<br />

phase space diagram. The complex order parameter of this swarm can be def<strong>in</strong>ed to be the magnitude and<br />

phase of the centroid of this swarm<br />

= 1 X<br />

<br />

(14.122)<br />

<br />

The centroid of the ensemble of po<strong>in</strong>ts on the phase diagram has a<br />

magnitude designat<strong>in</strong>g the offset of the centroid from the center of<br />

the circular phase diagram, and which is the phase of this centroid.<br />

A uniform distribution of po<strong>in</strong>ts around the unit circle will lead to a<br />

centroid =0. Correlated motion leads to a bunch<strong>in</strong>g of the po<strong>in</strong>ts<br />

around some phase value lead<strong>in</strong>g to a non-zero centroid and angle<br />

. If the swarm acts like a fully-coupled s<strong>in</strong>gle oscillator then ≈ 1<br />

with an appropriate phase .<br />

The Kuramoto model[Kur75, Str00] <strong>in</strong>corporates W<strong>in</strong>free’s<br />

<strong>in</strong>tuition by mapp<strong>in</strong>g the limit cycles onto a simple circular phase<br />

diagram and <strong>in</strong>corporat<strong>in</strong>g the long-term dynamics of coupled oscillators<br />

<strong>in</strong> terms of the relative phases for a mean-field system. That<br />

is, the angular velocity of the phase ˙ for the oscillator is<br />

˙ = +<br />

=1<br />

X<br />

Γ ( − ) (14.123)<br />

=1<br />

Figure 14.11: Order parameter for<br />

weakly-coupled oscillators.


390 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

Figure 14.12: Kuramoto model of collective synchronization of coupled oscillators. The left and center<br />

plots show the time and coupl<strong>in</strong>g strength dependence of the order parameter . Therightplotshowsthe<br />

frequency dependence <strong>in</strong>clud<strong>in</strong>g coupl<strong>in</strong>g (solid l<strong>in</strong>e) and without coupl<strong>in</strong>g (dashed l<strong>in</strong>e).<br />

where =1 2. Kuramoto recognized that mean-field coupl<strong>in</strong>g was the most tractable system to solve,<br />

that is, a system where the coupl<strong>in</strong>g is applicable equally to all the oscillators. Moreover, he assumed an<br />

equally-weighted, pure s<strong>in</strong>usoidal coupl<strong>in</strong>g for the coupl<strong>in</strong>g term Γ ( − ) between the coupled oscillators.<br />

That is, he assumed<br />

Γ ( − )= s<strong>in</strong>( − ) (14.124)<br />

where ≥ 0 is the coupl<strong>in</strong>g strength, and the factor 1 <br />

ensures that the model is well behaved as →∞.<br />

Kuramoto assumed that the frequency distribution () was unimodular and symmetric about the mean<br />

frequency Ω, thatis(Ω + ) =(Ω − ).<br />

This problem can be simplified by exploit<strong>in</strong>g the rotational symmetry and transform<strong>in</strong>g to a frame of<br />

reference that is rotat<strong>in</strong>g at an angular frequency Ω. That is, use the transformation = − Ω where<br />

is measured <strong>in</strong> the rotat<strong>in</strong>g frame. This makes () unimodular with a symmetric frequency distribution<br />

about =0. The phase velocity <strong>in</strong> this rotat<strong>in</strong>g frame is<br />

˙ = +<br />

X<br />

=1<br />

<br />

s<strong>in</strong>( − ) (14.125)<br />

Kuramoto observed that the phase-space distribution can be expressed <strong>in</strong> terms of the order parameters <br />

<strong>in</strong> that equation 14122 canbemultipliedonbothsidesby − <br />

to give<br />

(− ) = 1 <br />

X<br />

=1<br />

( − )<br />

(14.126)<br />

Equat<strong>in</strong>g the imag<strong>in</strong>ary parts yields<br />

s<strong>in</strong> ( − )= 1 <br />

X<br />

s<strong>in</strong> ( − ) (14.127)<br />

=1<br />

This allows equation 14125 to be written as<br />

˙ = + s<strong>in</strong>( − ) (14.128)<br />

for =1 2. Equation 14128 reflects the mean-field aspect of the model <strong>in</strong> that each oscillator is<br />

attracted to the phase of the mean field rather than to the phase of another <strong>in</strong>dividual oscillator.<br />

Simulations showed that the evolution of the order parameter with coupl<strong>in</strong>g strength is as illustrated<br />

<strong>in</strong> figure 1412. This simulation shows (1) for all when below a certa<strong>in</strong> threshold , the order parameter<br />

decays to an <strong>in</strong>coherent jitter as expected for random scatter of po<strong>in</strong>ts. (2) When this <strong>in</strong>coherent<br />

state becomes unstable and the order parameter grows exponentially reflect<strong>in</strong>g the nucleation of small<br />

clusters of oscillators that are mutually synchronized. (3) The population of <strong>in</strong>dividual oscillators splits<br />

<strong>in</strong>to two groups. The oscillators near the center of the distribution lock together <strong>in</strong> phase at the mean


14.12. COLLECTIVE SYNCHRONIZATION OF COUPLED OSCILLATORS 391<br />

angular frequency Ω and co-rotate with average phase (), whereas those frequencies ly<strong>in</strong>g further from<br />

the center cont<strong>in</strong>ue to rotate <strong>in</strong>dependently at their natural frequencies and drift relative to the coherent<br />

cluster frequency Ω. As a consequence this mixed state is only partially synchronized as illustrated on the<br />

right side of figure 1412. The synchronized fraction has a -function behavior for the frequency distribution<br />

which grows <strong>in</strong> <strong>in</strong>tensity with further <strong>in</strong>crease <strong>in</strong> . The unsynchronized component has nearly the orig<strong>in</strong>al<br />

frequency distribution () except that it is depleted <strong>in</strong> the region of the locked frequency due to strength<br />

absorbed by the -function component.<br />

Kuramoto’s toy model nicely illustrates the essential features of the evolution of collective synchronization<br />

with coupl<strong>in</strong>g strength. It has been applied to the study neuronal synchronization <strong>in</strong> the bra<strong>in</strong>[Cum07]. The<br />

model illustrates that the collective synchronization of coupled oscillators leads to a component that has a<br />

s<strong>in</strong>gle frequency for correlated motion which can be much narrower than the <strong>in</strong>herent frequency distribution<br />

of the ensemble of coupled oscillators.<br />

14.12 Example: Collective motion <strong>in</strong> nuclei<br />

The nucleus is an unusual quantal system that <strong>in</strong>volves the coupled motion of the many nucleons. It<br />

exhibits features characteristic of the many-body classical coupled oscillator with coupl<strong>in</strong>g between all the<br />

valence nucleons. Nuclear structure can be described by a shell model of <strong>in</strong>dividual nucleons bound <strong>in</strong> weakly<br />

<strong>in</strong>teract<strong>in</strong>g orbits <strong>in</strong> a central average mean field that is produced by the summed attraction of all the nucleons<br />

<strong>in</strong> the nucleus. However, nuclei also exhibit features characteristic of collective rotation and vibration of a<br />

quantal fluid. For example, beautiful rotational bands up to sp<strong>in</strong> over 60~ are observed <strong>in</strong> heavy nuclei. These<br />

rotational bands are similar to those observed <strong>in</strong> the rotational structure of diatomic molecules. Act<strong>in</strong>ide<br />

nuclei also can fission <strong>in</strong>to two large fragments which is another manifestation of collective motion.<br />

The essential general feature of weakly-coupled identical oscillators is illustrated by the solutions of the<br />

three l<strong>in</strong>early-coupled identical oscillators where the most symmetric state is displaced <strong>in</strong> frequency from the<br />

rema<strong>in</strong><strong>in</strong>g states For identical oscillators, one state is displaced significantly <strong>in</strong> energy from the rema<strong>in</strong><strong>in</strong>g<br />

− 1 degenerate states. This most symmetric state is pushed downwards <strong>in</strong> energy if the residual coupl<strong>in</strong>g<br />

force is attractive, and it is pushed upwards if the coupl<strong>in</strong>g force is repulsive. This symmetric state corresponds<br />

to the coherent oscillation of all the coupled oscillators, and carries all of the strength for the correspond<strong>in</strong>g<br />

dom<strong>in</strong>ant multipole for the coupl<strong>in</strong>g force. In the nucleus this state corresponds to coherent shape oscillations<br />

of many nucleons.<br />

The weak residual electric quadrupole and octupole nucleon-nucleon correlations <strong>in</strong> the nucleon-nucleon<br />

<strong>in</strong>teractions generate collective quadrupole and octupole motion <strong>in</strong> nuclei. The collective synchronization<br />

of such coherent quadrupole and octupole excitation leads to collective bands of states, that correspond to<br />

synchronized <strong>in</strong>-phase motion of the protons and neutrons <strong>in</strong> the valence oscillator shell. These modes<br />

correspond to rotations and vibrations about the center of mass. The attractive residual nucleon-nucleon<br />

<strong>in</strong>teraction couples the many <strong>in</strong>dividual particle excitations <strong>in</strong> a given shell produc<strong>in</strong>g one coherent state that<br />

is pushed downwards <strong>in</strong> energy far from the rema<strong>in</strong><strong>in</strong>g − 1 degenerate states. This coherent state <strong>in</strong>volves<br />

correlated motion of the nucleons that corresponds to a macroscopic oscillation of a charged fluid. For nonclosed<br />

shell nuclei like 238 U, the dom<strong>in</strong>ant quadrupole multipole <strong>in</strong> the residual nucleon-nucleon <strong>in</strong>teraction<br />

leads to the ground state be<strong>in</strong>g a coherent state correspond<strong>in</strong>g to ≈ 16 protons plus ≈ 20 neutrons oscillat<strong>in</strong>g<br />

<strong>in</strong> phase. The collective motion of the charged protons leads to electromagnetic 2 radiation with a transition<br />

decay amplitude be<strong>in</strong>g about 16 times larger than for a s<strong>in</strong>gle proton. This corresponds to radiative<br />

decay probability be<strong>in</strong>g enhanced by a factor of ≈ 256 relative to radiation by a s<strong>in</strong>gle proton. This collective<br />

state corresponds to a macroscopic quadrupole deformation at low excitation energies that exhibits both collective<br />

rotational and vibrational degrees of freedom. This coherent state is analogous to the correlated flow<br />

of <strong>in</strong>dividual water molecules <strong>in</strong> a tidal wave. The weaker octupole term <strong>in</strong> the residual <strong>in</strong>teraction leads to<br />

an octupole [pear-shaped] coupled oscillator coherent state ly<strong>in</strong>g slightly above the quadrupole coherent state.<br />

In contrast to the rotational motion of strongly-deformed quadrupole-deformed nuclei, the octupole deformation<br />

exhibits more vibrational-like properties than rotational motion of a charged tidal wave. Hamiltonian<br />

mechanics, based on the Routhian is used to make theoretical model calculations of the nuclear<br />

structure of 235 <strong>in</strong> the rotat<strong>in</strong>g body-fixed frame for comparison with the experimental data.


392 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

14.13 Summary<br />

This chapter has focussed on many—body coupled l<strong>in</strong>ear oscillator systems which are a ubiquitous feature <strong>in</strong><br />

nature. A summary of the ma<strong>in</strong> conclusions are the follow<strong>in</strong>g.<br />

Normal modes: It was shown that coupled l<strong>in</strong>ear oscillators exhibit normal modes and normal coord<strong>in</strong>ates<br />

that correspond to <strong>in</strong>dependent modes of oscillation with characteristic eigenfrequencies .<br />

General analytic theory for coupled l<strong>in</strong>ear oscillators Lagrangian mechanics was used to derive the<br />

general analytic procedure for solution of the many-body coupled oscillator problem which reduces to the<br />

conventional eigenvalue problem. A summary of the procedure for solv<strong>in</strong>g coupled oscillator problems is as<br />

follows:.<br />

1) Choose generalized coord<strong>in</strong>ates and evaluate and .<br />

= 1 2<br />

and<br />

= 1 2<br />

where the components of the T and V tensors are<br />

X<br />

˙ ˙ <br />

<br />

X<br />

<br />

<br />

(1441)<br />

(1442)<br />

≡<br />

X<br />

<br />

<br />

3X<br />

<br />

<br />

<br />

<br />

<br />

(1443)<br />

and<br />

µ 2 <br />

≡<br />

(1444)<br />

<br />

0<br />

2) Determ<strong>in</strong>e the eigenvalues us<strong>in</strong>g the secular determ<strong>in</strong>ant.<br />

11 − 2 11 12 − 2 12 13 − 2 13 <br />

12 − 2 12 22 − 2 22 23 − 2 23 <br />

13 − 2 13 23 − 2 23 33 − 2 33 <br />

=0 (1452)<br />

¯ ¯<br />

3) The eigenvectors are obta<strong>in</strong>ed by <strong>in</strong>sert<strong>in</strong>g the eigenvalues <strong>in</strong>to<br />

X ¡<br />

− 2 ¢<br />

=0 (1451)<br />

4) From the <strong>in</strong>itial conditions determ<strong>in</strong>e the complex scale factors where<br />

<br />

() ≡ <br />

(1458)<br />

5) Determ<strong>in</strong>e the normal coord<strong>in</strong>ates where each is a normal mode. The normal coord<strong>in</strong>ates can be<br />

expressed as<br />

η = {a} −1 q<br />

(1461)<br />

Few-body coupled oscillator systems The general analytic theory was used to determ<strong>in</strong>e the solutions<br />

for parallel and series coupl<strong>in</strong>gs of two and three l<strong>in</strong>ear oscillators. The phenomena observed <strong>in</strong>clude degenerate<br />

and non-degenerate eigenvalues and spurious center-of-mass oscillatory modes. There are two broad<br />

classifications for three or more coupled oscillators, that is, either complete coupl<strong>in</strong>g of all oscillators, or<br />

coupl<strong>in</strong>g of the nearest-neighbor oscillators. It is observed that the eigenvalue correspond<strong>in</strong>g to the most<br />

coherent motion of the coupled oscillators corresponds to the most collective motion and its eigenvalue is displaced<br />

the most <strong>in</strong> energy from the rema<strong>in</strong><strong>in</strong>g eigenvalues. For some systems this coherent collective mode


14.13. SUMMARY 393<br />

corresponded to a center-of-mass motion with no <strong>in</strong>ternal excitation of the other modes, while the other<br />

eigenvalues corresponded to modes with <strong>in</strong>ternal excitation of the oscillators such that the center of mass<br />

is stationary. The above procedure has been applied to two classification of coupl<strong>in</strong>g, complete coupl<strong>in</strong>g of<br />

many oscillators, and nearest neighbor coupl<strong>in</strong>g. Both degenerate and spurious center-of-mass modes were<br />

observed. Strong collective shape degrees of freedom <strong>in</strong> nuclei are examples of complete coupl<strong>in</strong>g due to the<br />

weak residual <strong>in</strong>teractions between nucleons <strong>in</strong> the nucleus. It was seen that, for many coupled oscillators,<br />

one coherent state separates from the other states and this coherent state carries the bulk of the collective<br />

strength.<br />

Discrete lattice cha<strong>in</strong> Transverse and longitud<strong>in</strong>al modes of motion on the discrete lattice cha<strong>in</strong> were discussed<br />

because of the important role it plays <strong>in</strong> nature, such as <strong>in</strong> crystall<strong>in</strong>e lattice structures. Both normal<br />

modes and travell<strong>in</strong>g waves were discussed <strong>in</strong>clud<strong>in</strong>g the phenomena of dispersion and cut-off frequencies.<br />

Molecules and the crystall<strong>in</strong>e lattice cha<strong>in</strong>s are examples where nearest neighbor coupl<strong>in</strong>g is manifest. It<br />

was shown that, for the −oscillator discrete lattice cha<strong>in</strong>, there are only <strong>in</strong>dependent longitud<strong>in</strong>al modes<br />

plus modes for the two transverse polarizations, and that the angular frequency ≤ 2 0 that is, a cut-off<br />

frequency exists.<br />

Damped coupled l<strong>in</strong>ear oscillators It was shown that l<strong>in</strong>early-damped coupled oscillator systems can<br />

be solved analytically us<strong>in</strong>g the concept of the Rayleigh dissipation function.<br />

Collective synchronization of coupled oscillators The Kuramoto schematic phase model was used<br />

to illustrate how weak residual forces can cause collective synchronization of the motion of many coupled<br />

oscillators. This is applicable to many large coupled systems such as nuclei, molecules, and biological systems.<br />

Problems<br />

1. Two particles, each with mass , move <strong>in</strong> one dimension <strong>in</strong> a region near a local m<strong>in</strong>imum of the potential<br />

energy where the potential energy is approximately given by<br />

where is a constant.<br />

(a) Determ<strong>in</strong>e the frequencies of oscillation.<br />

(b) Determ<strong>in</strong>e the normal coord<strong>in</strong>ates.<br />

2. What is degeneracy? When does it arise?<br />

= 1 2 (72 1 +4 2 2 +4 1 2 )<br />

3. The Lagrangian of three coupled oscillators is given by:<br />

3X<br />

=1<br />

∙ ˙<br />

2<br />

<br />

2<br />

F<strong>in</strong>d 2 () for the follow<strong>in</strong>g <strong>in</strong>itial conditions (at =0):<br />

¸<br />

− 2 <br />

+ 0 ( 1 2 + 2 3 )<br />

2<br />

( 1 2 3 )=( 0 0 0) :::::: ( ˙ 1 ˙ 2 ˙ 3 )=(0 0 0 )<br />

4. A mechanical analog of the benzene molecule comprises a discrete lattice cha<strong>in</strong> of 6 po<strong>in</strong>t masses connected<br />

<strong>in</strong> a plane hexagonal r<strong>in</strong>g by 6 identical spr<strong>in</strong>gs each with spr<strong>in</strong>g constant and length .<br />

a) List the wave numbers of the allowed undamped longitud<strong>in</strong>al stand<strong>in</strong>g waves.<br />

b) Calculate the phase velocity and group velocity for longitud<strong>in</strong>al travell<strong>in</strong>g waves on the r<strong>in</strong>g.<br />

c) Determ<strong>in</strong>e the time dependence of a longitud<strong>in</strong>al stand<strong>in</strong>g wave for a angular frequency =2 ,that<br />

is, twice the cut-off frequency.


394 CHAPTER 14. COUPLED LINEAR OSCILLATORS<br />

5. Consider a one dimensional, two-mass, three-spr<strong>in</strong>g system governed by the matrix ,<br />

µ 4 −2<br />

=<br />

−2 7<br />

such that = 2 ,<br />

(a) Determ<strong>in</strong>e the eigenfrequencies and normal coord<strong>in</strong>ates.<br />

(b) Choose a set of <strong>in</strong>itial conditions such that the system oscillates at its highest eigenfrequency.<br />

(c) Determ<strong>in</strong>e the solutions 1 () and 2 ().<br />

6. Four identical masses are connected by four identical spr<strong>in</strong>gs, spr<strong>in</strong>g constant and constra<strong>in</strong>ed to move<br />

on a frictionless circle of radius as shown on the left <strong>in</strong> the figure.<br />

a) How many normal modes of small oscillation are there?<br />

b) What are the eigenfrequencies of the small oscillations?<br />

c) Describe the motion of the four masses for each eigenfrequency.<br />

7. Consider the two identical coupled oscillators given on the right <strong>in</strong> the figure assum<strong>in</strong>g 1 = 2 = . Letboth<br />

oscillators be l<strong>in</strong>early damped with a damp<strong>in</strong>g constant . A force = 0 cos() is applied to mass 1 .<br />

Write down the pair of coupled differential equations that describe the motion. Obta<strong>in</strong> a solution by express<strong>in</strong>g<br />

the differential equations <strong>in</strong> terms of the normal coord<strong>in</strong>ates. Show that the normal coord<strong>in</strong>ates 1 and 2<br />

exhibit resonance peaks at the characteristic frequencies 1 and 2 respectively.<br />

8. As shown on the left below the mass moves horizontally along a frictionless rail. A pendulum is hung from<br />

with a weightless rod of length with a mass at its end.<br />

a) Prove that the eigenfrequencies are<br />

b) Describe the normal modes.<br />

1 =0 2 =<br />

r <br />

( + )<br />

<br />

x<br />

M


Chapter 15<br />

Advanced Hamiltonian mechanics<br />

15.1 Introduction<br />

This study of classical mechanics has <strong>in</strong>volved climb<strong>in</strong>g a vast mounta<strong>in</strong> of knowledge, while the pathway to<br />

the top has led us to elegant and beautiful theories that underlie much of modern physics. Be<strong>in</strong>g so close to<br />

the summit provides the opportunity to take a few extra steps <strong>in</strong> order to provide a glimpse of applications<br />

to physics at the summit. These are described <strong>in</strong> chapters 15 − 18.<br />

Hamilton’s development of Hamiltonian mechanics <strong>in</strong> 1834 is the crown<strong>in</strong>g achievement for apply<strong>in</strong>g variational<br />

pr<strong>in</strong>ciples to classical mechanics. A fundamental advantage of Hamiltonian mechanics is that it uses<br />

the conjugate coord<strong>in</strong>ates q p plus time , which is a considerable advantage <strong>in</strong> most branches of physics<br />

and eng<strong>in</strong>eer<strong>in</strong>g. Compared to Lagrangian mechanics, Hamiltonian mechanics has a significantly broader<br />

arsenal of powerful techniques that can be exploited to obta<strong>in</strong> an analytical solution of the <strong>in</strong>tegrals of the<br />

motion for complicated systems. In addition, Hamiltonian dynamics provides a means of determ<strong>in</strong><strong>in</strong>g the<br />

unknown variables for which the solution assumes a soluble form, and is ideal for study of the fundamental<br />

underly<strong>in</strong>g physics <strong>in</strong> applications to fields such as quantum or statistical physics. As a consequence, Hamiltonian<br />

mechanics has become the preem<strong>in</strong>ent variational approach used <strong>in</strong> modern physics. This chapter<br />

<strong>in</strong>troduces the follow<strong>in</strong>g four techniques <strong>in</strong> Hamiltonian mechanics: (1) the elegant Poisson bracket representation<br />

of Hamiltonian mechanics, which played a pivotal role <strong>in</strong> the development of quantum theory; (2)<br />

the powerful Hamilton-Jacobi theory coupled with Jacobi’s development of canonical transformation theory;<br />

(3) action-angle variable theory; and (4) canonical perturbation theory.<br />

Prior to further development of the theory of Hamiltonian mechanics, it is useful to summarize the major<br />

formula relevant to Hamiltonian mechanics that have been presented <strong>in</strong> chapters 7 8 and 9.<br />

Action functional :<br />

As discussed <strong>in</strong> chapter 92, Hamiltonian mechanics is built upon Hamilton’s action functional<br />

Z 2<br />

(q p) = (q ˙q) (15.1)<br />

1<br />

Hamilton’s Pr<strong>in</strong>ciple of least action states that<br />

Z 2<br />

(q p) = (q ˙q) =0 (15.2)<br />

1<br />

Generalized momentum :<br />

In chapter 72, the generalized (canonical) momentum was def<strong>in</strong>ed <strong>in</strong> terms of the Lagrangian to be<br />

(q ˙q)<br />

≡ (15.3)<br />

˙ <br />

Chapter 92 def<strong>in</strong>ed the generalized momentum <strong>in</strong> terms of the action functional to be<br />

=<br />

(q p)<br />

<br />

(15.4)<br />

395


396 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

Generalized energy (q ˙ ) :<br />

Jacobi’s Generalized Energy (q ˙ ) was def<strong>in</strong>ed <strong>in</strong> equation 737 as<br />

(q ˙q) ≡ X µ<br />

<br />

(q ˙q)<br />

˙ − (q ˙q) (15.5)<br />

˙<br />

<br />

<br />

Hamiltonian function:<br />

The Hamiltonian (q p) was def<strong>in</strong>ed <strong>in</strong> terms of the generalized energy (q ˙q) plus the generalized<br />

momentum. That is<br />

(q p) ≡ (q ˙q)= X <br />

˙ − (q ˙q)=p · ˙q−(q ˙q) (15.6)<br />

where p q correspond to -dimensional vectors, e.g. q ≡ ( 1 2 ) and the scalar product p· ˙q = P ˙ .<br />

Chapter 82 used a Legendre transformation to derive this relation between the Hamiltonian and Lagrangian<br />

functions. Note that whereas the Lagrangian (q ˙q) is expressed <strong>in</strong> terms of the coord<strong>in</strong>ates q plus<br />

conjugate velocities ˙q, the Hamiltonian (q p) is expressed <strong>in</strong> terms of the coord<strong>in</strong>ates q plus their<br />

conjugate momenta p. For scleronomic systems, plus assum<strong>in</strong>g the standard Lagrangian, then equations<br />

744 and 729 give that the Hamiltonian simplifies to equal the total mechanical energy, that is, = + .<br />

Generalized energy theorem:<br />

The equations of motion lead to the generalized energy theorem which states that the time dependence<br />

of the Hamiltonian is related to the time dependence of the Lagrangian.<br />

(q p)<br />

= X "<br />

#<br />

X<br />

˙ (q ˙q)<br />

+ (q) − (15.7)<br />

<br />

<br />

<br />

<br />

Note that if all the generalized non-potential forces and Lagrange multiplier terms are zero, and if the<br />

Lagrangian is not an explicit function of time, then the Hamiltonian is a constant of motion.<br />

Hamilton’s equations of motion:<br />

Chapter 83 showed that a Legendre transform plus the Lagrange-Euler equations led to Hamilton’s<br />

equations of motion. Hamilton derived these equations of motion directly from the action functional, as<br />

shown <strong>in</strong> chapter 92<br />

(q p)<br />

<br />

=1<br />

(q p)<br />

˙ = (15.8)<br />

<br />

"<br />

˙ = − <br />

<br />

#<br />

X <br />

(q p)+ + <br />

<br />

(15.9)<br />

<br />

(q ˙q)<br />

= −<br />

<br />

=1<br />

(15.10)<br />

Note the symmetry of Hamilton’s two canonical equations. The canonical variables are treated<br />

as <strong>in</strong>dependent canonical variables Lagrange was the first to derive the canonical equations but he did not<br />

recognize them as a basic set of equations of motion. Hamilton derived the canonical equations of motion<br />

from his fundamental variational pr<strong>in</strong>ciple and made them the basis for a far-reach<strong>in</strong>g theory of dynamics.<br />

Hamilton’s equations give 2 first-order differential equations for for each of the degrees of freedom.<br />

Lagrange’s equations give second-order differential equations for the variables ˙ <br />

Hamilton-Jacobi equation:<br />

Hamilton used Hamilton’s Pr<strong>in</strong>ciple to derive the Hamilton-Jacobi equation.<br />

<br />

<br />

+ (q p) =0 (15.11)<br />

The solution of Hamilton’s equations is trivial if the Hamiltonian is a constant of motion, or when a set<br />

of generalized coord<strong>in</strong>ate can be identified for which all the coord<strong>in</strong>ates are constant, or are cyclic (also<br />

called ignorable coord<strong>in</strong>ates). Jacobi developed the mathematical framework of canonical transformations<br />

required to exploit the Hamilton-Jacobi equation.


15.2. POISSON BRACKET REPRESENTATION OF HAMILTONIAN MECHANICS 397<br />

15.2 Poisson bracket representation of Hamiltonian mechanics<br />

15.2.1 Poisson Brackets<br />

Poisson brackets were developed by Poisson, who was a student of Lagrange. Hamilton’s canonical equations<br />

of motion describe the time evolution of the canonical variables ( ) <strong>in</strong> phase space. Jacobi showed that the<br />

framework of Hamiltonian mechanics can be restated <strong>in</strong> terms of the elegant and powerful Poisson bracket<br />

formalism. The Poisson bracket representation of Hamiltonian mechanics provides a direct l<strong>in</strong>k between<br />

classical mechanics and quantum mechanics.<br />

The Poisson bracket of any two cont<strong>in</strong>uous functions of generalized coord<strong>in</strong>ates ( ) and ( ) is<br />

def<strong>in</strong>ed to be<br />

[ ] <br />

≡ X µ <br />

− <br />

<br />

(15.12)<br />

<br />

<br />

Note that the above def<strong>in</strong>ition of the Poisson bracket leads to the follow<strong>in</strong>g identity, antisymmetry, l<strong>in</strong>earity,<br />

Leibniz rules, and Jacobi Identity.<br />

[ ] =0 (15.13)<br />

[ ] =− [ ] (15.14)<br />

[ + ]=[ ]+[ ] (15.15)<br />

[ ]=[ ] + [ ] (15.16)<br />

0=[ [ ]] + [ []] + [ [ ]] (15.17)<br />

where and are functions of the canonical variables plus time. Jacobi’s identity; (1517) states that<br />

the sum of the cyclic permutation of the double Poisson brackets of three functions is zero. Jacobi’s identity<br />

plays a useful role <strong>in</strong> Hamiltonian mechanics as will be shown.<br />

15.2.2 Fundamental Poisson brackets:<br />

The Poisson brackets of the canonical variables themselves are called the fundamental Poisson brackets.<br />

They are<br />

[ ] <br />

= X µ <br />

− <br />

<br />

= X ( · 0 − 0 · )=0<br />

<br />

<br />

<br />

(15.18)<br />

[ ] <br />

= X µ <br />

− <br />

<br />

= X (0 · − · 0) = 0<br />

<br />

<br />

<br />

(15.19)<br />

[ ] <br />

= X µ <br />

− <br />

<br />

= X ( · − 0 · 0) = <br />

<br />

<br />

<br />

(15.20)<br />

In summary, the fundamental Poisson brackets equal<br />

[ ] <br />

=0 (15.21)<br />

[ ] <br />

=0 (15.22)<br />

[ ] <br />

= − [ ] <br />

= (15.23)<br />

Note that the Poisson bracket is antisymmetric under <strong>in</strong>terchange <strong>in</strong> and It is <strong>in</strong>terest<strong>in</strong>g that the only<br />

non-zero fundamental Poisson bracket is for conjugate variables where = that is<br />

[ ] <br />

=1 (15.24)


398 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

15.2.3 Poisson bracket <strong>in</strong>variance to canonical transformations<br />

The Poisson brackets are <strong>in</strong>variant under a canonical transformation from one set of canonical variables<br />

( ) to a new set of canonical variables ( ) where → (q p) and → (q p). Thisisshown<br />

by transform<strong>in</strong>g equation 1512 to the new variables by the follow<strong>in</strong>g derivation<br />

[ ] <br />

= X µ <br />

− <br />

<br />

(15.25)<br />

<br />

<br />

= X µ µ <br />

+ <br />

<br />

− µ <br />

+ <br />

<br />

(15.26)<br />

<br />

<br />

The terms can be rearranged to give<br />

[ ] <br />

= X <br />

µ <br />

<br />

[ ] <br />

+ <br />

<br />

[ ] <br />

<br />

(15.27)<br />

Let = and replace by , and use the fact that the fundamental Poisson brackets [ ] <br />

=0<br />

and [ ] <br />

= ,thenequation1525 reduces to<br />

[ ] <br />

= X <br />

µ <br />

[ ]+ <br />

[ ] = X <br />

<br />

<br />

<br />

(15.28)<br />

That is<br />

Similarly<br />

[ ] <br />

= X <br />

[ ]=− <br />

<br />

(15.29)<br />

µ <br />

<br />

[ ] <br />

+ <br />

<br />

[ ] <br />

<br />

(15.30)<br />

lead<strong>in</strong>g to<br />

[ ] <br />

= <br />

(15.31)<br />

<br />

Substitut<strong>in</strong>g equations (1529) and (1531) <strong>in</strong>to equation (1527) gives<br />

[ ] <br />

= X µ <br />

− <br />

<br />

=[ ]<br />

<br />

(15.32)<br />

<br />

<br />

Thus the canonical variable subscripts ( ) and ( ) can be ignored s<strong>in</strong>ce the Poisson bracket is<br />

<strong>in</strong>variant to any canonical transformation of canonical variables. The counter argument is that if the Poisson<br />

bracket is <strong>in</strong>dependent of the transformation, then the transformation is canonical.<br />

15.1 Example: Check that a transformation is canonical<br />

The <strong>in</strong>dependence of Poisson brackets to canonical transformations can be used to test if a transformation<br />

is canonical. Assume that the transformation equations between two sets of coord<strong>in</strong>ates are given by<br />

=ln<br />

³1+ 1 2 cos ´<br />

=2<br />

³1+ 1 2 cos ´<br />

1 2 s<strong>in</strong> <br />

Evaluat<strong>in</strong>g the Poisson brackets gives [ ] =0, [ ] =0 while<br />

[ ] = <br />

− <br />

<br />

− 1 2 cos <br />

=<br />

1+ 1 2 cos [− s<strong>in</strong>2 +(1+ 1 1 1 2 s<strong>in</strong> 2 <br />

2 cos ) 2 cos ]+<br />

1+ 1 2 cos [cos +(1+ 1 2 cos )<br />

− 1 2 ]=1<br />

Therefore if are canonical with a Poisson bracket [ ] =1,thensoare s<strong>in</strong>ce [ ]=1=[ ]


15.2. POISSON BRACKET REPRESENTATION OF HAMILTONIAN MECHANICS 399<br />

S<strong>in</strong>ce it has been shown that this transformation is canonical, it is possible to go further and determ<strong>in</strong>e<br />

the function that generates this transformation. Solv<strong>in</strong>g the transformation equations for and give<br />

= ¡ − 1 ¢ 2<br />

sec 2 <br />

=2 ¡ − 1 ¢ tan <br />

S<strong>in</strong>ce the transformation is canonical, there exists a generat<strong>in</strong>g function 3 ( ) such that<br />

= − 3<br />

<br />

= − 3<br />

<br />

The transformation function 3 ( ) can be obta<strong>in</strong>ed us<strong>in</strong>g<br />

3 ( ) = 3<br />

+ 3<br />

= −− <br />

<br />

h ¡<br />

= −<br />

− 1 ¢ i 2<br />

tan − ¡ − 1 ¢ h<br />

2 ¡<br />

tan = − − 1 ¢ i<br />

2<br />

tan <br />

This then gives that the required generat<strong>in</strong>g function is<br />

3 ( ) = ¡ − 1 ¢ 2<br />

tan <br />

This example illustrates how to determ<strong>in</strong>e a useful generat<strong>in</strong>g function and prove that the transformation is<br />

canonical.<br />

15.2.4 Correspondence of the commutator and the Poisson Bracket<br />

In classical mechanics there is a formal correspondence between the Poisson bracket and the commutator.<br />

This can be shown by deriv<strong>in</strong>g the Poisson Bracket of four functions taken <strong>in</strong> two pairs. The derivation<br />

requires deriv<strong>in</strong>g the two possible Poisson Brackets <strong>in</strong>volv<strong>in</strong>g three functions.<br />

[ 1 2 ] = X ½µ µ ¾<br />

1 2 1 2 <br />

2 + 1 − 2 + 1<br />

<br />

<br />

= [ 1 ] 2 + 1 [ 2 ] (15.33)<br />

[ 1 2 ] = [ 1 ] 2 + 1 [ 2 ] (15.34)<br />

These two Poisson Brackets for three functions can be used to derive the Poisson Bracket of four functions,<br />

taken <strong>in</strong> pairs. This can be accomplished two ways us<strong>in</strong>g either equation 1533 or 1534<br />

[ 1 2 1 2 ] = [ 1 1 2 ] 2 + 1 [ 2 1 2 ]<br />

= {[ 1 1 ] 2 + 1 [ 1 2 ]} 2 + 1 {[ 2 1 ] 2 + 1 [ 2 2 ]}<br />

= [ 1 1 ] 2 2 + 1 [ 1 2 ] 2 + 1 [ 2 1 ] 2 + 1 1 [ 2 2 ] (15.35)<br />

The alternative approach gives<br />

[ 1 2 1 2 ] = [ 1 2 1 ] 2 + 1 [ 1 2 2 ]<br />

= [ 1 1 ] 2 2 + 1 [ 2 1 ] 2 + 1 [ 1 2 ] 2 + 1 1 [ 2 2 ] (15.36)<br />

These two alternate derivations give different relations for the same Poisson Bracket. Equat<strong>in</strong>g the alternative<br />

equations 1535 and 1536 gives that<br />

[ 1 1 ]( 2 2 − 2 2 )=( 1 1 − 1 1 )[ 2 2 ]<br />

This can be factored <strong>in</strong>to separate relations, the left-hand side for body 1 and the right-hand side for body<br />

2.<br />

( 1 1 − 1 1 )<br />

= ( 2 2 − 2 2 )<br />

= (15.37)<br />

[ 1 1 ] [ 2 2 ]


400 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

S<strong>in</strong>ce the left-hand ratio holds for 1 1 <strong>in</strong>dependent of 2 2 , and vise versa, then they must equal<br />

a constant that does not depend on 1 1 does not depend on 2 2 , and must commute with<br />

( 1 1 − 1 1 ).Thatis, must be a constant number <strong>in</strong>dependent of these variables.<br />

( 1 1 − 1 1 )= [ 1 1 ] ≡ X µ 1 1<br />

− <br />

1 1<br />

(15.38)<br />

<br />

<br />

Equation 1538 is an especially important result which states that to with<strong>in</strong> a multiplicative constant number<br />

, there is a one-to-one correspondence between the Poisson Bracket and the commutator of two <strong>in</strong>dependent<br />

functions. An important implication is that if two functions, have a Poisson Bracket that is zero, then<br />

the commutator of the two functions also must be zero, that is, and commute.<br />

Consider the special case where the variables 1 and 1 correspond to the fundamental canonical variables,<br />

( ). Then the commutators of the fundamental canonical variables are given by<br />

− = [ ]= (15.39)<br />

− = [ ]=0 (15.40)<br />

− = [ ]=0 (15.41)<br />

In 1925, Paul Dirac, a 23-year old graduate student at Bristol, recognized that the formal correspondence<br />

between the Poisson bracket <strong>in</strong> classical mechanics, and the correspond<strong>in</strong>g commutator, provides a logical<br />

and consistent way to bridge the chasm between the Hamiltonian formulation of classical mechanics, and<br />

quantum mechanics. He realized that mak<strong>in</strong>g the assumption that the constant ≡ ~, leads to Heisenberg’s<br />

fundamental commutation relations <strong>in</strong> quantum mechanics, as is discussed <strong>in</strong> chapter 1831. Assum<strong>in</strong>g that<br />

≡ ~ provides a logical and consistent way that builds quantization directly <strong>in</strong>to classical mechanics, rather<br />

than us<strong>in</strong>g ad-hoc, case-dependent, hypotheses as was used by the older quantum theory of Bohr.<br />

15.2.5 Observables <strong>in</strong> Hamiltonian mechanics<br />

Poisson brackets, and the correspond<strong>in</strong>g commutation relations, are especially useful for elucidat<strong>in</strong>g which<br />

observables are constants of motion, and whether any two observables can be measured simultaneously and<br />

exactly. The properties of any observable are determ<strong>in</strong>ed by the follow<strong>in</strong>g two criteria.<br />

Time dependence:<br />

The total time differential of a function ( ) is def<strong>in</strong>ed by<br />

<br />

= <br />

+ X <br />

µ <br />

<br />

˙ + <br />

<br />

˙ <br />

<br />

(15.42)<br />

Hamilton’s canonical equations give that<br />

Substitut<strong>in</strong>g these <strong>in</strong> the above relation gives<br />

<br />

= <br />

+ X <br />

˙ = <br />

<br />

(15.43)<br />

˙ = − <br />

<br />

(15.44)<br />

µ <br />

− <br />

<br />

<br />

that is<br />

<br />

= +[ ] (15.45)<br />

<br />

This important equation states that the total time derivative of any function ( ) can be expressed <strong>in</strong><br />

terms of the partial time derivative plus the Poisson bracket of ( ) with the Hamiltonian.


15.2. POISSON BRACKET REPRESENTATION OF HAMILTONIAN MECHANICS 401<br />

Any observable ( ) will be a constant of motion if <br />

<br />

=0, and thus equation (1545) gives<br />

<br />

<br />

+[ ] =0<br />

(If is a constant of motion)<br />

That is, it is a constant of motion when<br />

<br />

=[ ] (15.46)<br />

<br />

Moreover, this can be extended further to the statement that if the constant of motion is not explicitly<br />

time dependent then<br />

[ ] =0 (15.47)<br />

The Poisson bracket with the Hamiltonian is zero for a constant of motion that is not explicitly time<br />

dependent. Often it is more useful to turn this statement around with the statement that if [ ] =0 and<br />

<br />

=0 then =0,imply<strong>in</strong>gthat is a constant of motion.<br />

<br />

<br />

<br />

Independence<br />

Consider two observables ( ) and ( ). The <strong>in</strong>dependence of these two observables is determ<strong>in</strong>ed<br />

by the Poisson bracket<br />

[ ] =− [ ] (15.48)<br />

If this Poisson bracket is zero, that is, if the two observables ( ) and ( ) commute, then their<br />

values are <strong>in</strong>dependent and can be measured <strong>in</strong>dependently. However, if the Poisson bracket [ ] 6= 0,that<br />

is ( ) and ( ) do not commute, then and are correlated s<strong>in</strong>ce <strong>in</strong>terchang<strong>in</strong>g the order of<br />

the Poisson bracket changes the sign which implies that the measured value for depends on whether is<br />

simultaneously measured.<br />

A useful property of Poisson brackets is that if and both are constants of motion, then the double<br />

Poisson bracket [ [ ]] = 0. This can be proved us<strong>in</strong>g Jacobi’s identity<br />

[ [ ]] + [ [ ]] + [ [ ]] = 0 (15.49)<br />

If [ ] =0and [ ] =0 then [ [ ]] = 0 that is, the Poisson bracket [ ] commutes with . Note<br />

that if and do not depend explicitly on time, that is <br />

<br />

= <br />

<br />

=0, then comb<strong>in</strong><strong>in</strong>g equations (1545)<br />

and (1549) leads to Poisson’s Theorem that relates the total time derivatives.<br />

∙<br />

<br />

¸ ∙ <br />

[ ] = + ¸<br />

(15.50)<br />

<br />

This implies that if and are <strong>in</strong>variants, that is <br />

<br />

<strong>in</strong>variant if and are not explicitly time dependent.<br />

= <br />

<br />

=0 then the Poisson bracket [ ] is an<br />

15.2 Example: Angular momentum:<br />

Angular momentum, provides an example of the use of Poisson brackets to elucidate which observables<br />

can be determ<strong>in</strong>ed simultaneously. Consider that the Hamiltonian is time <strong>in</strong>dependent with a spherically<br />

symmetric potential (). Then it is best to treat such a spherically symmetric potential us<strong>in</strong>g spherical<br />

coord<strong>in</strong>ates s<strong>in</strong>ce the Hamiltonian is <strong>in</strong>dependent of both and .<br />

The Poisson Brackets <strong>in</strong> classical mechanics can be used to tell us if two observables will commute. S<strong>in</strong>ce<br />

() is time <strong>in</strong>dependent, then the Hamiltonian <strong>in</strong> spherical coord<strong>in</strong>ates is<br />

Ã<br />

!<br />

= + = 1 2 + 2 <br />

2 2 +<br />

2 <br />

2 s<strong>in</strong> 2 + ()<br />

<br />

Evaluate the Poisson bracket us<strong>in</strong>g the above Hamiltonian gives<br />

[ ]=0


402 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

S<strong>in</strong>ce is not an explicit function of time, <br />

<br />

=0 then <br />

<br />

=0 that is, the angular momentum about<br />

the axis = is a constant of motion.<br />

The Poisson bracket of the total angular momentum 2 commutes with the Hamiltonian, that is<br />

£<br />

2 ¤ "<br />

#<br />

= 2 +<br />

2 <br />

s<strong>in</strong> 2 =0<br />

2 <br />

s<strong>in</strong> 2 <br />

S<strong>in</strong>ce the total angular momentum 2 = 2 + is not explicitly time dependent, then it also must be<br />

a constant of motion. Note that Noether’s theorem gives that both the angular momenta 2 and are<br />

constants of motion. Also s<strong>in</strong>ce the Poisson brackets are<br />

[ ] = 0<br />

£<br />

2 ¤ = 0<br />

then Jacobi’s identity, equation 1517 can be used to imply that<br />

[ £ 2 <br />

¤<br />

]=0<br />

That is, the Poisson bracket £ ¤<br />

£ ¤<br />

2 is a constant of motion. Note that if 2 and commute, that is,<br />

£ 2 <br />

¤ =0 then they can be measured simultaneously with unlimited accuracy, and this also satisfies that<br />

2 commutes with .<br />

The ( ) components of the angular momentum are given by<br />

X<br />

X<br />

= (r × p) <br />

= ( − )<br />

=<br />

=<br />

=1<br />

X<br />

(r × p) <br />

=<br />

=1<br />

X<br />

(r × p) <br />

=<br />

=1<br />

Evaluate the Poisson bracket<br />

[ ] =<br />

X<br />

∙µ <br />

− µ<br />

<br />

+<br />

<br />

=1 <br />

X<br />

= [(0) + (0) + ( − )] = <br />

=1<br />

Similarly, Poisson brackets for are<br />

=1<br />

X<br />

( − )<br />

=1<br />

X<br />

( − )<br />

=1<br />

[ ] = <br />

[ ] = <br />

[ ] = <br />

<br />

− µ<br />

<br />

+<br />

<br />

where and are taken <strong>in</strong> a right-handed cyclic order. This usually is written <strong>in</strong> the form<br />

[ ]= <br />

<br />

<br />

− <br />

<br />

<br />

<br />

¸<br />

where the Levi-Civita density equals zero if two of the <strong>in</strong>dices are identical, otherwise it is +1 for a<br />

cyclic permutation of , and −1 for a non-cyclic permutation.<br />

Note that s<strong>in</strong>ce these Poisson brackets are nonzero, the components of the angular momentum <br />

do not commute and thus simultaneously they cannot be measured precisely. Thus we see that although 2 and<br />

are simultaneous constants of motion, where the subscript can be either or only one component<br />

can be measured simultaneously with 2 . This behavior is exhibited by rigid-body rotation where the body<br />

precesses around one component of the total angular momentum, , such that the total angular momentum,<br />

2 , plus the component along one axis, are constants of motion. Then 2 + 2 = 2 − 2 is constant<br />

but not the <strong>in</strong>dividual or .


15.2. POISSON BRACKET REPRESENTATION OF HAMILTONIAN MECHANICS 403<br />

15.2.6 Hamilton’s equations of motion<br />

An especially important application of Poisson brackets is that Hamilton’s canonical equations of motion<br />

can be expressed directly <strong>in</strong> the Poisson bracket form. The Poisson bracket representation of Hamiltonian<br />

mechanics has important implications to quantum mechanics as will be described <strong>in</strong> chapter 18.<br />

In equation (1545) assume that is a fundamental coord<strong>in</strong>ate, that is, ≡ .S<strong>in</strong>ce is not explicitly<br />

time dependent, then<br />

<br />

<br />

= <br />

+[ ] (15.51)<br />

= 0+ X µ <br />

− <br />

<br />

<br />

<br />

= X µ<br />

<br />

− 0 · <br />

<br />

<br />

<br />

= <br />

(15.52)<br />

<br />

That is<br />

˙ =[ ]= <br />

(15.53)<br />

<br />

Similarly consider the fundamental canonical momentum ≡ . S<strong>in</strong>ce it is not explicitly time dependent,<br />

then<br />

<br />

<br />

= <br />

+[ ] (15.54)<br />

= 0+ X µ <br />

− <br />

<br />

<br />

<br />

= X µ<br />

0 · − · <br />

<br />

<br />

<br />

= − <br />

(15.55)<br />

<br />

That is<br />

˙ =[ ]=− <br />

(15.56)<br />

<br />

Thus, it is seen that the Poisson bracket form of the equations of motion <strong>in</strong>cludes the Hamilton equations<br />

of motion. That is,<br />

˙ = [ ]= <br />

<br />

(15.57)<br />

˙ = [ ]=− <br />

<br />

(15.58)<br />

The above shows that the full structure of Hamilton’s equations of motion can be expressed directly <strong>in</strong><br />

terms of Poisson brackets.<br />

The elegant formulation of Poisson brackets has the same form <strong>in</strong> all canonical coord<strong>in</strong>ates as the Hamiltonian<br />

formulation. However, the normal Hamilton canonical equations <strong>in</strong> classical mechanics assume implicitly<br />

that one can specify the exact position and momentum of a particle simultaneously at any po<strong>in</strong>t <strong>in</strong> time<br />

which is applicable only to classical mechanics variables that are cont<strong>in</strong>uous functions of the coord<strong>in</strong>ates,<br />

and not to quantized systems. The important feature of the Poisson Bracket representation of Hamilton’s<br />

equations is that it generalizes Hamilton’s equations <strong>in</strong>to a form (1557 1558) where the Poisson bracket is<br />

equally consistent with both classical and quantum mechanics <strong>in</strong> that it allows for non-commut<strong>in</strong>g canonical<br />

variables and Heisenberg’s Uncerta<strong>in</strong>ty Pr<strong>in</strong>ciple. Thus the generalization of Hamilton’s equations, via use<br />

of the Poisson brackets, provides one of the most powerful analytic tools applicable to both classical and<br />

quantal dynamics. It played a pivotal role <strong>in</strong> derivation of quantum theory as described <strong>in</strong> chapter 18.


404 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

15.3 Example: Lorentz force <strong>in</strong> electromagnetism<br />

Consider a charge and mass <strong>in</strong> a constant electromagnetic fields with scalar potential Φ and vector<br />

potential Chapter 610 showed that the Lagrangian for electromagnetism can be written as<br />

The generalized momentum then is given by<br />

= 1 ẋ · ẋ−(Φ − A · ẋ)<br />

2<br />

Thus the Hamiltonian can be written as<br />

p = =ẋ + A<br />

ẋ<br />

=(p · ẋ) − = (p−A)2<br />

2<br />

The Hamilton equations of motion give<br />

+ Φ<br />

and<br />

ẋ=[x]= (p−A)<br />

<br />

ṗ =[p] =−∇Φ + {(p−A) × (∇ × A)}<br />

<br />

Def<strong>in</strong>e the magnetic fieldtobe<br />

and the electric fieldtobe<br />

then the Lorentz force can be written as<br />

B ≡ ∇ × A<br />

E = − ∇Φ − A<br />

<br />

F = ṗ = (E + ẋ × B)<br />

15.4 Example: Wavemotion:<br />

Assume that one is deal<strong>in</strong>g with travel<strong>in</strong>g waves of the form Ψ = ( 1 −) for a one-dimensional<br />

conservative system of many identical coupled l<strong>in</strong>ear oscillators. Then evaluat<strong>in</strong>g the follow<strong>in</strong>g Poisson<br />

brackets gives<br />

[ ] = 0<br />

[ ] = 0<br />

[ ] = 0<br />

[ ] = 0<br />

Thus and are constants of motion. However,<br />

[ ] 6= 0<br />

[ ] 6= 0<br />

Thus one cannot simultaneously measure the conjugate variables ( ) or ( ). This is the Uncerta<strong>in</strong>ty<br />

Pr<strong>in</strong>ciple that is manifest by all forms of wave motion <strong>in</strong> classical and quantal mechanics as discussed <strong>in</strong><br />

chapter 3113


15.2. POISSON BRACKET REPRESENTATION OF HAMILTONIAN MECHANICS 405<br />

15.5 Example: Two-dimensional, anisotropic, l<strong>in</strong>ear oscillator<br />

Consider a mass bound by an anisotropic, two-dimensional, l<strong>in</strong>ear oscillator potential. As discussed<br />

<strong>in</strong> chapter 11 , the motion can be described as ly<strong>in</strong>g entirely <strong>in</strong> the − plane that is perpendicular to the<br />

angular momentum . It is <strong>in</strong>terest<strong>in</strong>g to derive the equations of motion for this system us<strong>in</strong>g the Poisson<br />

bracket representation of Hamiltonian mechanics.<br />

The k<strong>in</strong>etic energy is given by<br />

( ˙ ˙) = 1 2 ¡ ˙ 2 + ˙ 2¢<br />

The l<strong>in</strong>ear b<strong>in</strong>d<strong>in</strong>g is reproduced assum<strong>in</strong>g a quadratic scalar potential energy of the form<br />

( ) = 1 2 ¡ 2 + 2¢ + <br />

where is the anharmonic strength that coupled the modes of the isotropic l<strong>in</strong>ear oscillator.<br />

a) NORMAL MODES: As discussed <strong>in</strong> chapter 14 , a transformation to the normal modes of the system<br />

is given by us<strong>in</strong>g variables ( ) where ≡ √ 1<br />

2<br />

( + ) and ≡ √ 1 2<br />

( − ), that is<br />

≡ 1 √<br />

2<br />

( + ) ≡ 1 √<br />

2<br />

( − )<br />

Express the k<strong>in</strong>etic and potential energies <strong>in</strong> terms of the new coord<strong>in</strong>ates gives<br />

( ˙ ˙) = 1 ∙ ³<br />

4 ˙ + ˙<br />

´2 ³ ´2¸<br />

+ ˙ − ˙ = 1 ³<br />

2´<br />

2 ˙ 2 + ˙<br />

= 1 h<br />

4 ( + ) 2 +( − ) 2i + 1 2 ¡ 2 − 2¢ = 1 2 ( + ) 2 + 1 ( − ) 2<br />

2<br />

Note that the coord<strong>in</strong>ate transformation makes the Lagrangian separable, that is<br />

= 1 ³<br />

2 ˙ 2 + ˙ 2´ − 1 2 ( + ) 2 + 1 2 ( − ) 2 = + <br />

where<br />

= 1 2 ˙2 − 1 ( + ) 2<br />

<br />

2<br />

= 1 2 ˙ 2 − 1 ( − ) 2<br />

2<br />

This shows that that the transformation has separated the system <strong>in</strong>to two normal modes that are harmonic<br />

oscillators with angular frequencies<br />

r r<br />

+ <br />

− <br />

1 =<br />

2 =<br />

<br />

<br />

Note that the non-isotropic harmonic oscillator reduces to the isotropic l<strong>in</strong>ear oscillator when =0.<br />

b) HAMILTONIAN: The canonical momenta are given by<br />

= <br />

˙ = ˙<br />

= = ˙<br />

˙<br />

The def<strong>in</strong>ition of the Hamiltonian gives<br />

1 ¡<br />

= ˙ + ˙ − = <br />

2<br />

2 + 2 1<br />

¢<br />

+<br />

2 ( + ) 2 + 1 ( − ) 2<br />

2<br />

Note that this can be factored as<br />

where<br />

= + <br />

= 1<br />

2 2 + 1 2 ( + ) 2 = 1<br />

2 2 + 1 ( − ) 2<br />

2


406 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

Us<strong>in</strong>g the Poisson Bracket expression for the time dependence, equation 1545 and us<strong>in</strong>g the fact that<br />

the Hamiltonian is not explicitly time dependent, that is, <br />

<br />

=0,gives<br />

<br />

<br />

= <br />

<br />

= <br />

<br />

+[ ]=0+[ + ]=[ ]<br />

<br />

<br />

+ <br />

<br />

<br />

<br />

− <br />

<br />

<br />

− <br />

<br />

<br />

=0<br />

Similarly <br />

<br />

=0. This implies that the Hamiltonians for both normal modes, and are time<strong>in</strong>dependent<br />

constants of motion which are equal to the total energy for each mode.<br />

c) ANGULAR MOMENTUM: The angular momentum for motion <strong>in</strong> the plane is perpendicular to<br />

the plane with a magnitude of<br />

= ( − )<br />

The time dependence of the angular momentum is given by<br />

<br />

<br />

= <br />

<br />

+[ ] =0+ − <br />

+ <br />

− <br />

<br />

= + + − − + =2<br />

Note that if =0 then the two eigenfrequencies, are degenerate, = , that is, the system reduces to<br />

the isotropic harmonic oscillator <strong>in</strong> the plane that was discussed <strong>in</strong> chapter 119. In addition, <br />

=0<br />

for =0 that is, the angular momentum <strong>in</strong> the plane is a constant of motion when =0.<br />

d) SYMMETRY TENSOR: The symmetry tensor was def<strong>in</strong>ed <strong>in</strong> chapter 1193 to be<br />

0 = <br />

2 + 1 2 <br />

where and can correspond to either or . The symmetry tensor def<strong>in</strong>es the orientation of the major<br />

axis of the elliptical orbit for the two-dimensional, isotropic, l<strong>in</strong>ear oscillator as described <strong>in</strong> chapter 1193<br />

The isotropic oscillator has been shown to have two normal modes that are degenerate, therefore and<br />

are equally good normal modes. The Hamiltonian showed that, for =0 the Hamiltonian gives that the<br />

total energy is conserved, as well as the energies for each of the two normal modes which are.<br />

= 2 <br />

2 + 1 2 2<br />

Consider the matrix element<br />

0 = <br />

2 + 1 2 <br />

where each can represent or . Then for each matrix element<br />

= 2 <br />

2 + 1 2 2<br />

0 <br />

<br />

= 0 <br />

<br />

+[ ]=0+ 0 <br />

<br />

<br />

− 0 <br />

+ 0 <br />

− 0 <br />

=0<br />

That is, each matrix element 0 12 commutes with the Hamiltonian<br />

£<br />

<br />

0<br />

¤ =0<br />

Thus the Poisson Brackets representation of Hamiltonian mechanics has been used to prove that the<br />

symmetry tensor 0 = <br />

2 + 1 2 is a constant of motion for the isotropic harmonic oscillator. That is,<br />

all the elements 0 , 0 and 0 of the symmetric tensor A0 commute with the Hamiltonian.<br />

Note that the three constants of motion, L, A 0 and H for the isotropic, two-dimensional, l<strong>in</strong>ear oscillator<br />

form a closed algebra under the Poisson Bracket formalism.<br />

15.6 Example: The eccentricity vector<br />

Chapter 1184 showed that Hamilton’s eccentricity vector for the <strong>in</strong>verse square-law attractive force,<br />

A ≡ (p × L)+(ˆr)


15.2. POISSON BRACKET REPRESENTATION OF HAMILTONIAN MECHANICS 407<br />

is a constant of motion that specifies the major axis of the elliptical orbit. The eccentricity vector for the<br />

<strong>in</strong>verse-square-law force can be <strong>in</strong>vestigated us<strong>in</strong>g Poisson Brackets as was done for the symmetry tensor<br />

above. It can be shown that<br />

[ ] = <br />

µ p<br />

2<br />

[ ] = −2<br />

2 + <br />

<br />

<br />

(a)<br />

Note that the bracket on the right-hand side of equation () equals the Hamiltonian for the <strong>in</strong>verse squarelaw<br />

attractive force, and thus the Poisson bracket equals<br />

µ p<br />

2<br />

[ ]=−2<br />

2 + <br />

= −2 <br />

<br />

For the Hamiltonian it can be shown that the Poisson bracket<br />

[ A] =0<br />

That is, the eccentricity vector commutes with the Hamiltonian and thus it is a constant of motion. Previously<br />

this result was obta<strong>in</strong>ed directly us<strong>in</strong>g the equations of motion as given <strong>in</strong> equation 1187. Note that the three<br />

constants of motion, L, A and H form a closed algebra under the Poisson Bracket formalism similar to<br />

the triad of constants of motion, L, A 0 and H that occur for the two-dimensional, isotropic l<strong>in</strong>ear oscillator<br />

described above. Examples 155 and 156 illustrate that the Poisson Brackets representation of Hamiltonian<br />

mechanics is a powerful probe of the underly<strong>in</strong>g physics, as well as confirm<strong>in</strong>g the results obta<strong>in</strong>ed directly<br />

from the equations of motion as described <strong>in</strong> chapter 1184 and 11 9 3 .<br />

15.2.7 Liouville’s Theorem<br />

Liouvilles Theorem illustrates an application of Poisson Brackets<br />

to Hamiltonian phase space that has important implications for<br />

statistical physics. The trajectory of a s<strong>in</strong>gle particle <strong>in</strong> phase<br />

space is completely determ<strong>in</strong>ed by the equations of motion if the<br />

<strong>in</strong>itial conditions are known. However, many-body systems have<br />

so many degrees of freedom it becomes impractical to solve all<br />

the equations of motion of the many bodies. An example is a<br />

statistical ensemble <strong>in</strong> a gas, a plasma, or a beam of particles.<br />

Usually it is not possible to specify the exact po<strong>in</strong>t <strong>in</strong> phase space<br />

for such complicated systems. However, it is possible to def<strong>in</strong>e an<br />

ensemble of po<strong>in</strong>ts <strong>in</strong> phase space that encompasses all possible<br />

trajectories for the complicated system. That is, the statistical<br />

distribution of particles <strong>in</strong> phase space can be specified.<br />

Consider a density of representative po<strong>in</strong>ts <strong>in</strong> (q p) phase<br />

space. The number of systems <strong>in</strong> the volume element is<br />

= (15.59)<br />

p<br />

i<br />

dp<br />

i<br />

q<br />

i<br />

p<br />

dq<br />

i<br />

i<br />

q<br />

i<br />

where it is assumed that the <strong>in</strong>f<strong>in</strong>itessimal volume element<br />

= 1 2 1 2 conta<strong>in</strong>s many possible systems<br />

so that can be considered a cont<strong>in</strong>uous distribution. For <strong>in</strong> phase space<br />

Figure 15.1: Inf<strong>in</strong>itessimal element of area<br />

the conjugate variables ( ) shown <strong>in</strong> figure 151, thenumber<br />

of representative po<strong>in</strong>ts mov<strong>in</strong>g across the left-hand edge <strong>in</strong>to<br />

the area per unit time is<br />

˙ (15.60)<br />

Thenumberofrepresentativepo<strong>in</strong>tsflow<strong>in</strong>g out of the area along the right-hand edge is<br />

∙<br />

˙ + ( ˙ ) ¸<br />

(15.61)


408 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

Hence the net <strong>in</strong>crease <strong>in</strong> <strong>in</strong> the <strong>in</strong>f<strong>in</strong>itessimal rectangular element due to flow <strong>in</strong> the horizontal<br />

direction is<br />

− ( ˙ ) (15.62)<br />

<br />

Similarly, the net ga<strong>in</strong> due to flow <strong>in</strong> the vertical direction is<br />

− ( ˙ ) (15.63)<br />

<br />

Thus the total <strong>in</strong>crease <strong>in</strong> the element per unit time is therefore<br />

∙ <br />

− ( ˙ )+ ¸<br />

( ˙ ) (15.64)<br />

<br />

Assume that the total number of po<strong>in</strong>ts must be conserved, then the total <strong>in</strong>crease <strong>in</strong> the number of<br />

po<strong>in</strong>ts <strong>in</strong>side the element must equal the net changes <strong>in</strong> on the <strong>in</strong>f<strong>in</strong>itessimal surface element per<br />

unit time. That is<br />

µ <br />

(15.65)<br />

<br />

Thus summ<strong>in</strong>g over all possible values of gives<br />

<br />

+ X <br />

∙ <br />

( ˙ )+ ¸<br />

( ˙ )<br />

<br />

=0 (15.66)<br />

or<br />

<br />

+ X ∙<br />

¸<br />

<br />

˙ + ˙ + X ∙ ˙<br />

+ ˙ ¸<br />

<br />

=0 (15.67)<br />

<br />

<br />

<br />

Insert<strong>in</strong>g Hamilton’s canonical equations <strong>in</strong>to both brackets and differentiat<strong>in</strong>g the last bracket results <strong>in</strong><br />

<br />

+ X ∙ <br />

− ¸<br />

<br />

+ X ∙ 2 ¸<br />

<br />

−<br />

2 <br />

=0 (15.68)<br />

<br />

<br />

<br />

The two terms <strong>in</strong> the last bracket cancel and thus<br />

<br />

+ X <br />

∙ <br />

− ¸<br />

<br />

= +[ ] =0 (15.69)<br />

<br />

However, this just equals <br />

, therefore<br />

<br />

= +[ ] =0 (15.70)<br />

<br />

This is called Liouville’s theorem which states that the rate of change of density of representative<br />

po<strong>in</strong>ts vanishes, that is, the density of po<strong>in</strong>ts is a constant <strong>in</strong> the Hamiltonian phase space along a specific<br />

trajectory. Liouville’s theorem means that the system acts like an <strong>in</strong>compressible fluid that moves such as to<br />

occupy an equal volume <strong>in</strong> phase space at every <strong>in</strong>stant, even though the shape of the phase-space volume<br />

may change, that is, the phase-space density of the fluid rema<strong>in</strong>s constant. Equation (1570) is another<br />

illustration of the basic Poisson bracket relation (1545) and the usefulness of Poisson brackets <strong>in</strong> physics.<br />

Liouville’s theorem is crucially important to statistical mechanics of ensembles where the exact knowledge<br />

of the system is unknown, only statistical averages are known. An example is <strong>in</strong> focuss<strong>in</strong>g of beams of charged<br />

particles by beam handl<strong>in</strong>g systems. At a focus of the beam, the transverse width <strong>in</strong> is m<strong>in</strong>imized, while<br />

the width <strong>in</strong> is largest s<strong>in</strong>ce the beam is converg<strong>in</strong>g to the focus, whereas a parallel beam has maximum<br />

width and m<strong>in</strong>imum spread<strong>in</strong>g width . However, the product rema<strong>in</strong>s constant throughout the<br />

focuss<strong>in</strong>g system. For a two dimensional beam, this applies equally for the and coord<strong>in</strong>ates, etc. It is<br />

obvious that the f<strong>in</strong>al beam quality for any beam transport system is ultimately limited by the emittance of<br />

the source of the beam, that is, the <strong>in</strong>itial area of the phase space distribution. Note that Liouville’s theorem<br />

only applies to Hamiltonian − phase space, not to − ˙ Lagrangian state space. As a consequence,<br />

Hamiltonian dynamics, rather than Lagrange dynamics, is used to discuss ensembles <strong>in</strong> statistical physics.<br />

Note that Liouville’s theorem is applicable only for conservative systems, that is, where Hamilton’s<br />

equations of motion apply. For dissipative systems the phase space volume shr<strong>in</strong>ks with time rather than<br />

be<strong>in</strong>g a constant of the motion.


15.3. CANONICAL TRANSFORMATIONS IN HAMILTONIAN MECHANICS 409<br />

15.3 Canonical transformations <strong>in</strong> Hamiltonian mechanics<br />

Hamiltonian mechanics is an especially elegant and powerful way to derive the equations of motion for complicated<br />

systems. Unfortunately, <strong>in</strong>tegrat<strong>in</strong>g the equations of motion to derive a solution can be a challenge.<br />

Hamilton recognized this difficulty, so he proposed us<strong>in</strong>g generat<strong>in</strong>g functions to make canonical transformations<br />

which transform the equations <strong>in</strong>to a known soluble form. Jacobi, a contemporary mathematician,<br />

recognized the importance of Hamilton’s pioneer<strong>in</strong>g developments <strong>in</strong> Hamiltonian mechanics, and therefore<br />

he developed a sophisticated mathematical framework for exploit<strong>in</strong>g the generat<strong>in</strong>g function formalism <strong>in</strong><br />

order to make the canonical transformations required to solve Hamilton’s equations of motion.<br />

In the Lagrange formulation, transform<strong>in</strong>g coord<strong>in</strong>ates ( ˙ ) to cyclic generalized coord<strong>in</strong>ates ( ˙ ),<br />

simplifies f<strong>in</strong>d<strong>in</strong>g the Euler-Lagrange equations of motion. For the Hamiltonian formulation, the concept of<br />

coord<strong>in</strong>ate transformations is extended to <strong>in</strong>clude simultaneous canonical transformation of both the spatial<br />

coord<strong>in</strong>ates and the conjugate momenta from ( ) to ( ), where both of the canonical variables<br />

are treated equally <strong>in</strong> the transformation. Compared to Lagrangian mechanics, Hamiltonian mechanics has<br />

twice as many variables which is an asset, rather than a liability, s<strong>in</strong>ce it widens the realm of possible<br />

canonical transformations.<br />

Hamiltonian mechanics has the advantage that generat<strong>in</strong>g functions can be exploited to make canonical<br />

transformations to f<strong>in</strong>d solutions, which avoids hav<strong>in</strong>g to use direct <strong>in</strong>tegration. Canonical transformations<br />

are the foundation of Hamiltonian mechanics; they underlie Hamilton-Jacobi theory and action-angle variable<br />

theory, both of which are powerful means for exploit<strong>in</strong>g Hamiltonian mechanics to solve problems <strong>in</strong> physics<br />

and eng<strong>in</strong>eer<strong>in</strong>g. The concept underly<strong>in</strong>g canonical transformations is that, if the equations of motion are<br />

simplified by us<strong>in</strong>g a new set of generalized variables (Q P) compared to us<strong>in</strong>g the orig<strong>in</strong>al set of variables<br />

(q p) then an advantage has been ga<strong>in</strong>ed. The solution, expressed <strong>in</strong> terms of the generalized variables<br />

(Q P) can be transformed back to express the solution <strong>in</strong> terms of the orig<strong>in</strong>al coord<strong>in</strong>ates, (q p).<br />

Only a specialized subset of transformations will be considered, namely canonical transformations that<br />

preserve the canonical form of Hamilton’s equations of motion. That is, given that the orig<strong>in</strong>al set of variables<br />

( ) satisfy Hamilton’s equations<br />

˙q =<br />

(q p)<br />

p<br />

− ṗ =<br />

(q p)<br />

q<br />

(15.71)<br />

for some Hamiltonian (q p) then the transformation to coord<strong>in</strong>ates ( ) ( ) is canonical<br />

if, and only if, there exists a function H(Q P) such that the P and Q are still governed by Hamilton’s<br />

equations. That is,<br />

˙Q =<br />

H(Q P)<br />

P<br />

− Ṗ =<br />

H(Q P)<br />

Q<br />

(15.72)<br />

where H(Q P) plays the role of the Hamiltonian for the new variables. Note that H(Q P) may be<br />

very different from the old Hamiltonian (q p). The <strong>in</strong>variance of the Poisson bracket to canonical<br />

transformations, chapter 1523, provides a powerful test that the transformation is canonical.<br />

Hamilton’s Pr<strong>in</strong>ciple of least action, discussed <strong>in</strong> chapter 9, statesthat<br />

Z 2<br />

Z 2<br />

= (q ˙q) = [p · ˙q − (q p)] =0 (15.73)<br />

1<br />

1<br />

Similarly, apply<strong>in</strong>g Hamilton’s Pr<strong>in</strong>ciple of least action to the new Lagrangian L(Q ˙Q) gives<br />

= <br />

Z 2<br />

1<br />

L(Q ˙Q) = <br />

Z 2<br />

1<br />

h<br />

i<br />

P · ˙Q − H(Q P) =0 (15.74)<br />

The discussion of gauge-<strong>in</strong>variant Lagrangians, chapter 93 showed that and L can be related by the total<br />

time derivative of a generat<strong>in</strong>g function where<br />

<br />

= L − (15.75)<br />

The generat<strong>in</strong>g function can be any well-behaved function with cont<strong>in</strong>uous second derivatives of both the<br />

old and new canonical variables p q P Q and Thus the <strong>in</strong>tegrands of (1573) and (1574) are related by<br />

p · ˙q − (q p)=<br />

h<br />

P · ˙Q − H(Q P)<br />

i<br />

+ <br />

<br />

(15.76)


410 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

where is a possible scale transformation. A scale transformation, such as chang<strong>in</strong>g units, is trivial, and will<br />

be assumed to be absorbed <strong>in</strong>to the coord<strong>in</strong>ates, mak<strong>in</strong>g =1 Assum<strong>in</strong>g that 6= 1is called an extended<br />

canonical transformation.<br />

15.3.1 Generat<strong>in</strong>g functions<br />

The generat<strong>in</strong>g function has to be chosen such that the transformation from the <strong>in</strong>itial variables (q p)<br />

to the f<strong>in</strong>al variables (Q P) is a canonical transformation. The chosen generat<strong>in</strong>g function contributes to<br />

(1576) only if it is a function of the old plus new variables. The four possible types of generat<strong>in</strong>g functions<br />

of the first k<strong>in</strong>d, are 1 (q Q), 2 (q P), 3 (p Q) ,and 4 (p P). These four generat<strong>in</strong>g functions<br />

lead to relatively simple canonical transformations, are shown below.<br />

Type 1: = 1 (q Q) :<br />

The total time derivative of the generat<strong>in</strong>g function = 1 (q Q) is given by<br />

(q Q)<br />

<br />

∙ 1 (q Q)<br />

=<br />

q<br />

· ˙q + 1(q Q)<br />

Q<br />

¸<br />

· ˙Q + 1(q Q)<br />

<br />

(15.77)<br />

Insert equation (1577) <strong>in</strong>to equation (1576), and assume that the trivial scale factor =1 then<br />

∙<br />

p − ¸<br />

∙<br />

1(q Q)<br />

· ˙q − (q p)= P + ¸<br />

1(q Q)<br />

· ˙Q − H(Q P)+ 1(q Q)<br />

q<br />

Q<br />

<br />

Assume that the generat<strong>in</strong>g function 1 determ<strong>in</strong>es the canonical variables p and P to be<br />

p = 1(q Q)<br />

q<br />

P = − 1(q Q)<br />

Q<br />

(15.78)<br />

then the terms <strong>in</strong> each square bracket cancel, lead<strong>in</strong>g to the required canonical transformation<br />

H(Q P)=(q p)+ 1(q Q)<br />

<br />

(15.79)<br />

Type 2: = 2 (q P) − Q · P :<br />

The total time derivative of the generat<strong>in</strong>g function = 2 (q P)−Q · P is given by<br />

∙<br />

<br />

= 2 (q P)<br />

· ˙q + ¸<br />

2(q P)<br />

· Ṗ − P · ˙Q − Ṗ · Q + 2(q P)<br />

q<br />

P<br />

<br />

(15.80)<br />

Insert this <strong>in</strong>to equation (1576) and assume that the trivial scale factor =1 then<br />

µ<br />

p − <br />

∙ ¸<br />

2(q P)<br />

2 (q P)<br />

· ˙q − (q p)=P · ˙Q − P · ˙Q+<br />

− Q · Ṗ − H(Q P)+ 2(q P)<br />

q<br />

P<br />

<br />

Assume that the generat<strong>in</strong>g function 2 determ<strong>in</strong>es the canonical variables p and Q to be<br />

p = 2(q P)<br />

q<br />

Q = 2(q P)<br />

P<br />

(15.81)<br />

then the terms <strong>in</strong> brackets cancel, lead<strong>in</strong>g to the required transformation<br />

H(Q P)=(q p)+ 2(q P)<br />

<br />

(15.82)


15.3. CANONICAL TRANSFORMATIONS IN HAMILTONIAN MECHANICS 411<br />

Type 3: = 3 (p Q)+q · p :<br />

The total time derivative of the generat<strong>in</strong>g function = 3 (p Q t)+q · p is given by<br />

∙<br />

<br />

= 3 (p Q)<br />

· ṗ + ¸<br />

3(p Q)<br />

· ˙Q + ˙q · p + q · ṗ + 3(p Q)<br />

p<br />

Q<br />

<br />

(15.83)<br />

Insert this <strong>in</strong>to equation (1576) and assume that the trivial scale factor =1 then<br />

∙<br />

− q+ ¸<br />

∙<br />

3(p Q)<br />

· ṗ − (q p)= P+ ¸<br />

3(p Q)<br />

· ˙Q − H(Q P)+ 3(p Q)<br />

p<br />

Q<br />

<br />

Assume that the generat<strong>in</strong>g function 3 determ<strong>in</strong>es the canonical variables q and P to be<br />

q = − 3(p Q)<br />

p<br />

P = − 3(p Q)<br />

Q<br />

(15.84)<br />

then the terms <strong>in</strong> brackets cancel, lead<strong>in</strong>g to the required transformation<br />

H(Q P)=(q p)+ 3(p Q)<br />

<br />

(15.85)<br />

Type 4: = 4 (p P)+q · p − Q · P :<br />

The total time derivative of the generat<strong>in</strong>g function = 4 (p P)+q · p − Q · P is given by<br />

∙<br />

<br />

= 4 (p P)<br />

· ṗ + ¸<br />

4(p P)<br />

· Ṗ + ˙q · p + q · ṗ − ˙Q · P − Q · Ṗ + 4(p P)<br />

p<br />

P<br />

<br />

(15.86)<br />

Insert this <strong>in</strong>to equation (1576) and assume that the trivial scale factor =1 then<br />

∙<br />

− q+ ¸<br />

∙ ¸<br />

4(p P)<br />

4 (p P)<br />

· ṗ − (q p)=<br />

− Q ·Ṗ − H(Q P)+ 4(p P)<br />

p<br />

P<br />

<br />

Assume that the generat<strong>in</strong>g function 4 determ<strong>in</strong>es the canonical variables q and Q to be<br />

q = − 4(p P)<br />

p<br />

Q = 4(p P)<br />

P<br />

(15.87)<br />

then the terms <strong>in</strong> brackets cancel, lead<strong>in</strong>g to the required transformation<br />

H(Q P)=(q p)+ 4(p P)<br />

(15.88)<br />

<br />

Note that the last three generat<strong>in</strong>g functions require the <strong>in</strong>clusion of additional bil<strong>in</strong>ear products of<br />

<strong>in</strong> order for the terms to cancel to give the required result. The addition of the bil<strong>in</strong>ear terms,<br />

ensures that the resultant generat<strong>in</strong>g function is the same us<strong>in</strong>g any of the four generat<strong>in</strong>g functions<br />

1 2 3 4 . Frequently the 2 (q P) generat<strong>in</strong>g function is the most convenient. The four possible<br />

generat<strong>in</strong>g functions of the first k<strong>in</strong>d, given above, are related by Legendre transformations. A canonical<br />

transformation does not have to conform to only one of the four generat<strong>in</strong>g functions for all the degrees<br />

of freedom, they can be a mixture of different flavors for the different degrees of freedom. The properties of<br />

the generat<strong>in</strong>g functions are summarized <strong>in</strong> table 151.<br />

Table 151 Canonical transformation generat<strong>in</strong>g functions<br />

Generat<strong>in</strong>g function Generat<strong>in</strong>g function derivatives Trivial special examples<br />

= 1 (q Q)<br />

= 1<br />

<br />

= − 1<br />

<br />

1 = = = − <br />

= 2 (q P) − Q · P = 2<br />

<br />

= 2<br />

<br />

2 = = = <br />

= 3 (p Q)+q · p = − 3<br />

<br />

= − 3<br />

<br />

3 = = − = − <br />

= 4 (p P)+q · p − Q · P = − 4<br />

<br />

= 4<br />

<br />

4 = = = −


412 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

The partial derivatives of the generat<strong>in</strong>g functions determ<strong>in</strong>e the correspond<strong>in</strong>g conjugate variables<br />

not explicitly <strong>in</strong>cluded <strong>in</strong> the generat<strong>in</strong>g function . Note that, for the first trivial example 1 = the<br />

old momenta become the new coord<strong>in</strong>ates, = and vice versa, = − . This illustrates that it is<br />

better to name them “conjugate variables” rather than “momenta” and “coord<strong>in</strong>ates”.<br />

In summary, Jacobi has developed a mathematical framework for f<strong>in</strong>d<strong>in</strong>g the generat<strong>in</strong>g function <br />

required to make a canonical transformation to a new Hamiltonian H(Q P), that has a known solution.<br />

That is,<br />

H(Q P)=(q p)+ <br />

(15.89)<br />

<br />

When H(Q P) is a constant, then a solution has been obta<strong>in</strong>ed. The <strong>in</strong>verse transformation for this solution<br />

Q() P() → q() p() now can be used to express the f<strong>in</strong>al solution <strong>in</strong> terms of the orig<strong>in</strong>al variables of the<br />

system.<br />

Note the special case when H(Q P)=0 then equation 1589 has been reduced to the Hamilton-Jacobi<br />

relation (1511)<br />

(q p)+ <br />

=0<br />

(1511)<br />

In this case, the generat<strong>in</strong>g function determ<strong>in</strong>es the action functional required to solve the Hamilton-<br />

Jacobi equation (15110). S<strong>in</strong>ce equation (1589) has transformed the Hamiltonian (q p) → H(Q P)<br />

for which H(Q P)=0, then the solution Q() P() for the Hamiltonian H(Q P)=0is obta<strong>in</strong>ed easily.<br />

This approach underlies Hamilton-Jacobi theory presented <strong>in</strong> chapter 154<br />

15.3.2 Applications of canonical transformations<br />

The canonical transformation procedure may appear unnecessarily complicated for solv<strong>in</strong>g the examples<br />

given <strong>in</strong> this book, but it is essential for solv<strong>in</strong>g the complicated systems that occur <strong>in</strong> nature. For example,<br />

canonical transformations can be used to transform time-dependent, (non-autonomous) Hamiltonians to<br />

time-<strong>in</strong>dependent, (autonomous) Hamiltonians for which the solutions are known. Example 1519 describes<br />

such a system. Canonical transformations provide a remarkably powerful approach for solv<strong>in</strong>g the equations<br />

of motion <strong>in</strong> Hamiltonian mechanics, especially when us<strong>in</strong>g the Hamilton-Jacobi approach discussed <strong>in</strong><br />

chapter 154.<br />

15.7 Example: The identity canonical transformation<br />

The identity transformation 2 (q P) = q · P satisfies (1589) if the follow<strong>in</strong>g relations are satisfied<br />

= 2<br />

<br />

= , = 2<br />

<br />

= , H=. Note that the new and old coord<strong>in</strong>ates are identical, hence 2 = <br />

generates the identity transformation = = .<br />

15.8 Example: The po<strong>in</strong>t canonical transformation<br />

Consider the po<strong>in</strong>t transformation 2 (q · P) = (q)·P where (q) is some function of q.<br />

transformation satisfies (1589) if the follow<strong>in</strong>g relations are satisfied = 2<br />

<br />

= ( ), = 2<br />

<br />

H=. Po<strong>in</strong>t transformations correspond to po<strong>in</strong>t-to-po<strong>in</strong>t transformations of coord<strong>in</strong>ates.<br />

15.9 Example: The exchange canonical transformation<br />

This<br />

<br />

= ()<br />

<br />

The identity transformation 1 (q Q) =q · Q satisfies (1589) if the follow<strong>in</strong>g relations are satisfied<br />

= 1<br />

<br />

= , = − 1<br />

<br />

= − , H= That is, the coord<strong>in</strong>ates and momenta have been <strong>in</strong>terchanged.<br />

15.10 Example: Inf<strong>in</strong>itessimal po<strong>in</strong>t canonical transformation<br />

Consider an <strong>in</strong>f<strong>in</strong>itessimal po<strong>in</strong>t canonical transformation, that is <strong>in</strong>f<strong>in</strong>itesimally close to a po<strong>in</strong>t identity.<br />

satisfies (1589) if the follow<strong>in</strong>g relations are satisfied<br />

2 (q · P) =q · P+(q P)<br />

= 2<br />

<br />

= + <br />

(q P)


15.3. CANONICAL TRANSFORMATIONS IN HAMILTONIAN MECHANICS 413<br />

= 2<br />

<br />

Thus the <strong>in</strong>f<strong>in</strong>itessimal changes <strong>in</strong> and are given by<br />

(q P)<br />

= + <br />

<br />

(q P)<br />

(q p) = − = <br />

<br />

(q p) = − = −<br />

(q P)<br />

<br />

(q P)<br />

= <br />

= −<br />

+ ( 2 )<br />

<br />

(q P)<br />

+ ( 2 )<br />

<br />

Thus (q P) is the generator of the <strong>in</strong>f<strong>in</strong>itessimal canonical transformation.<br />

15.11 Example: 1-D harmonic oscillator via a canonical transformation<br />

The classic one-dimensional harmonic oscillator provides an example of the use of canonical transformations.<br />

Consider the Hamiltonian where 2 = then<br />

= 2<br />

2 + 2<br />

2 = 1 ¡<br />

2 + 2 2 2¢<br />

2<br />

This form of the Hamiltonian is a sum of two squares suggest<strong>in</strong>g a canonical transformation for which<br />

is cyclic <strong>in</strong> a new coord<strong>in</strong>ate. A guess for a canonical transformation is of the form = cot which<br />

is of the 1 (q Q) type where 1 equals 1 ( ) = 2<br />

2<br />

cot Us<strong>in</strong>g (1578) gives<br />

Solv<strong>in</strong>gforthecoord<strong>in</strong>ates( ) yields<br />

= 1( )<br />

<br />

= − 1( )<br />

<br />

= cot <br />

= 2<br />

2<br />

s<strong>in</strong> 2 <br />

r<br />

2<br />

=<br />

s<strong>in</strong> <br />

(a)<br />

= √ 2 cos (b)<br />

Insert<strong>in</strong>g these <strong>in</strong>to gives<br />

H =(cos 2 +s<strong>in</strong> 2 )=<br />

which implies that is a cyclic coord<strong>in</strong>ate.<br />

The Hamiltonian is conservative, s<strong>in</strong>ce it does not explicitly depend on time, and it equals the total energy<br />

s<strong>in</strong>ce the transformation to generalized coord<strong>in</strong>ates is time <strong>in</strong>dependent. Thus<br />

H = = <br />

S<strong>in</strong>ce<br />

˙ = H<br />

= <br />

then<br />

= + <br />

Substitut<strong>in</strong>g <strong>in</strong>to () gives the well known solution of the one-dimensional harmonic oscillator<br />

r<br />

2<br />

= s<strong>in</strong>( + )<br />

2


414 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

15.4 Hamilton-Jacobi theory<br />

Hamilton used the Pr<strong>in</strong>ciple of Least Action to derive the Hamilton-Jacobi relation (chapter 153)<br />

(q p)+ <br />

=0<br />

(1511)<br />

where q p refer to the 1 ≤ ≤ variables and ( ( 1 ) 1 ( 2 ) 2 ) is the action functional. Integration<br />

of this first-order partial differential equation is non trivial which is a major handicap for practical<br />

exploitation of the Hamilton-Jacobi equation. This stimulated Jacobi to develop the mathematical framework<br />

for canonical transformation that are required to solve the Hamilton-Jacobi equation. Jacobi’s approach<br />

is to exploit generat<strong>in</strong>g functions for mak<strong>in</strong>g a canonical transformation to a new Hamiltonian H(Q P)<br />

that equals zero.<br />

H(Q P)=(q p)+ =0 (15.90)<br />

<br />

The generat<strong>in</strong>g function for solv<strong>in</strong>g the Hamilton-Jacobi equation then equals the action functional .<br />

The Hamilton-Jacobi theory is based on select<strong>in</strong>g a canonical transformation to new coord<strong>in</strong>ates ( )<br />

all of which are either constant, or the are cyclic, which implies that the correspond<strong>in</strong>g momenta are<br />

constants. In either case, a solution to the equations of motion is obta<strong>in</strong>ed. A remarkable feature of Hamilton-<br />

Jacobi theory is that the canonical transformation is completely characterized by a s<strong>in</strong>gle generat<strong>in</strong>g function,<br />

. The canonical equations likewise are characterized by a s<strong>in</strong>gle Hamiltonian function, . Moreover, the<br />

generat<strong>in</strong>g function and Hamiltonian function are l<strong>in</strong>ked together by equation 1511 The underly<strong>in</strong>g<br />

goal of Hamilton-Jacobi theory is to transform the Hamiltonian to a known form such that the canonical<br />

equations become directly <strong>in</strong>tegrable. S<strong>in</strong>ce this transformation depends on a s<strong>in</strong>gle scalar function, the<br />

problem is reduced to solv<strong>in</strong>g a s<strong>in</strong>gle partial differential equation.<br />

15.4.1 Time-dependent Hamiltonian<br />

Jacobi’s complete <strong>in</strong>tegral ( )<br />

The pr<strong>in</strong>ciple underly<strong>in</strong>g Jacobi’s approach to Hamilton-Jacobi theory is to provide a recipe for f<strong>in</strong>d<strong>in</strong>g<br />

the generat<strong>in</strong>g function = needed to transform the Hamiltonian (q p) to the new Hamiltonian<br />

H(Q P) us<strong>in</strong>g equation 1590 When the derivatives of the transformed Hamiltonian H(Q P) are zero,<br />

then the equations of motion become<br />

˙ = H =0 (15.91)<br />

<br />

˙<br />

= − H =0 (15.92)<br />

<br />

and thus and are constants of motion. The new Hamiltonian H must be related to the orig<strong>in</strong>al<br />

Hamiltonian by a canonical transformation for which<br />

H(Q P) = (q p)+ <br />

(15.93)<br />

<br />

Equations 1591 and 1592 are automatically satisfied if the new Hamiltonian H =0s<strong>in</strong>ce then equation<br />

1593 gives that the generat<strong>in</strong>g function satisfies equation 1590<br />

Any of the four types of generat<strong>in</strong>g function can be used. Jacobi chose the type 2 generat<strong>in</strong>g function<br />

as be<strong>in</strong>g the most useful for many practical cases, that is, ( ) which is called Jacobi’s complete<br />

<strong>in</strong>tegral.<br />

For generat<strong>in</strong>g functions 1 and 2 the generalized momenta are derived from the action by the derivative<br />

= <br />

(154)<br />

<br />

Use this generalized momentum to replace <strong>in</strong> the Hamiltonian , given <strong>in</strong> equation (1593) leads to the<br />

Hamilton-Jacobi equation expressed <strong>in</strong> terms of the action <br />

( 1 ; ; )+ =0 (15.94)<br />

1


15.4. HAMILTON-JACOBI THEORY 415<br />

The Hamilton-Jacobi equation, (1594) can be written more compactly us<strong>in</strong>g tensors q and ∇ to designate<br />

( 1 ) and <br />

1<br />

<br />

<br />

respectively. That is<br />

(q ∇ )+ =0 (15.95)<br />

<br />

Equation (1595) is a first-order partial differential equation <strong>in</strong> +1 variables which are the old spatial<br />

coord<strong>in</strong>ates plus time . The new momenta have not been specified except that they are constants<br />

s<strong>in</strong>ce H =0<br />

Assume the existence of a solution of (1595) of the form ( )=( 1 ; 1 +1 ; ) where<br />

the generalized momenta = 1 2 plus are the +1<strong>in</strong>dependent constants of <strong>in</strong>tegration <strong>in</strong> the<br />

transformed frame. One constant of <strong>in</strong>tegration is irrelevant to the solution s<strong>in</strong>ce only partial derivatives of<br />

( ) with respect to and are <strong>in</strong>volved. Thus, if is a solution of the first-order partial differential<br />

equation, then so is + where is a constant. Thus it can be assumed that one of the +1 constants of<br />

<strong>in</strong>tegration is just an additive constant which can be ignored lead<strong>in</strong>g effectively to a solution<br />

( )=( 1 ; 1 ; ) (15.96)<br />

wherenoneofthe <strong>in</strong>dependent constants are solely additive. Such generat<strong>in</strong>g function solutions are called<br />

complete solutions of the first-order partial differential equations s<strong>in</strong>ce all constants of <strong>in</strong>tegration are known.<br />

It is possible to assume that the generalized momenta, are constants ,wherethe are the<br />

constants This allows the generalized momentum to be written as<br />

(q α)<br />

= (15.97)<br />

<br />

Similarly, Hamilton’s equations of motion give the conjugate coord<strong>in</strong>ate Q = β where are constants That<br />

is<br />

(q α)<br />

= = (15.98)<br />

<br />

The above procedure has determ<strong>in</strong>ed the complete set of 2 constants (Q = β P = α). It is possible to<br />

<strong>in</strong>vert the canonical transformation to express the above solution, which is expressed <strong>in</strong> terms of = <br />

and = back to the orig<strong>in</strong>al coord<strong>in</strong>ates, that is, = ( ) and momenta = ( ) which is<br />

the required solution.<br />

Hamilton’s pr<strong>in</strong>ciple function (q ; q )<br />

Hamilton’s approach to solv<strong>in</strong>g the Hamilton-Jacobi equation (1595) is to seek a canonical transformation<br />

from variables (p q) at time to a new set of constant quantities, which may be the <strong>in</strong>itial values (q 0 p 0 )<br />

at time = 0 Hamilton’s pr<strong>in</strong>ciple function ( ; ) is the generat<strong>in</strong>g function for this canonical<br />

transformation from the variables (q p) at time to the <strong>in</strong>itial variables (q 0 p 0 ) at time 0 . Hamilton’s<br />

pr<strong>in</strong>ciple function ( ; ) is directly related to Jacobi’s complete <strong>in</strong>tegral ( ).<br />

Note that is the generat<strong>in</strong>g function of a canonical transformation from the present time (q p)<br />

variables to the <strong>in</strong>itial (q 0 p 0 0 ), whereas Jacobi’s is the generat<strong>in</strong>g function of a canonical transformation<br />

from the present (q p) variables to the constant variables (Q = β P = α). For the Hamilton approach,<br />

the canonical transformation can be accomplished <strong>in</strong> two steps us<strong>in</strong>g by first transform<strong>in</strong>g from (q p)<br />

at time , to(β α), then transform<strong>in</strong>g from (β α) to (q 0 p 0 0 ) That is, this two-step process corresponds<br />

to<br />

(q; q )=(q α) − (q 0 α 0 ) (15.99)<br />

Hamilton’s pr<strong>in</strong>ciple function (q; q ) is related to Jacobi’s complete <strong>in</strong>tegral (q α) and it will not<br />

be discussed further <strong>in</strong> this book.


416 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

15.4.2 Time-<strong>in</strong>dependent Hamiltonian<br />

Frequently the Hamiltonian does not explicitly depend on time. For the standard Lagrangian with time<strong>in</strong>dependent<br />

constra<strong>in</strong>ts and transformation, then (q p) = which is the total energy. For this case,<br />

the Hamilton-Jacobi equation simplifies to give<br />

<br />

= −(q p)=− (α) (15.100)<br />

<br />

The <strong>in</strong>tegration of the time dependence is trivial, and thus the action <strong>in</strong>tegral for a time-<strong>in</strong>dependent Hamiltonian<br />

equals<br />

(q α) = (q α) − (α) (15.101)<br />

That is, the action <strong>in</strong>tegral has separated <strong>in</strong>to a time <strong>in</strong>dependent term (q α) which is called Hamilton’s<br />

characteristic function plus a time-dependent term − (α) . Thus us<strong>in</strong>g equations 1597 15101 gives<br />

that the generalized momentum is<br />

(q α)<br />

= (15.102)<br />

<br />

The physical significance of Hamilton’s characteristic function (q α) can be understood by tak<strong>in</strong>g the<br />

total time derivative<br />

Tak<strong>in</strong>gthetime<strong>in</strong>tegralthengives<br />

<br />

<br />

= X <br />

(q α)<br />

<br />

˙ = X <br />

˙ <br />

(q α) =Z X<br />

˙ =<br />

Z X<br />

(15.103)<br />

Note that this equals the abbreviated action described <strong>in</strong> chapter 923, thatis (q α) = 0 (q α)<br />

Insert<strong>in</strong>g the action (q α) <strong>in</strong>to the Hamilton-Jacobi equation (1512) gives<br />

(q α)<br />

(q; )= (α)<br />

<br />

(15.104)<br />

This is called the time-<strong>in</strong>dependent Hamilton-Jacobi equation. Usually it is convenient to have <br />

equal the total energy. However, sometimes it is more convenient to exclude the energy ( ) <strong>in</strong> the<br />

set, <strong>in</strong> which case = ( 1 2 −1 ); the Routhian exploits this feature.<br />

The equations of the canonical transformation expressed <strong>in</strong> terms of (q α) are<br />

(q α)<br />

= <br />

+ (α) (q α)<br />

=<br />

<br />

<br />

(15.105)<br />

These equations show that Hamilton’s characteristic function (q α) is itself the generat<strong>in</strong>g function of a<br />

time-<strong>in</strong>dependent canonical transformation from the old variables ( ) to a set of new variables<br />

= + (α) <br />

<br />

= (15.106)<br />

Table 152 summarizes the time-dependent and time-<strong>in</strong>dependent forms of the Hamilton-Jacobi equation.<br />

Table 152; Hamilton-Jacobi formulations<br />

Hamiltonian Time dependent ( ) Time <strong>in</strong>dependent ( )<br />

Transformed Hamiltonian H= 0 H is cyclic<br />

Canonical transformed variables All are constants of motion All are constants of motion<br />

Transformed equations of motion ˙ = H<br />

<br />

=0 therefore = <br />

˙ = H<br />

<br />

= therefore = + <br />

˙<br />

= − H<br />

<br />

=0 therefore = ˙<br />

= − H<br />

<br />

=0 therefore = <br />

Generat<strong>in</strong>g function Jacobi’s complete <strong>in</strong>tegral (q P) Characteristic Function (q P)<br />

Hamilton-Jacobi equation<br />

Transformation equations<br />

( 1<br />

; <br />

1<br />

<br />

<br />

; )+ <br />

=0<br />

= <br />

<br />

= <br />

<br />

= <br />

( 1 ; <br />

1<br />

<br />

<br />

)=<br />

= <br />

<br />

= <br />

<br />

= +


15.4. HAMILTON-JACOBI THEORY 417<br />

15.4.3 Separation of variables<br />

Exploitation of the Hamilton-Jacobi theory requires f<strong>in</strong>d<strong>in</strong>g a suitable action function . When the Hamiltonian<br />

is time <strong>in</strong>dependent, then equation 15101 shows that the time dependence of the action <strong>in</strong>tegral<br />

separates out from the dependence on the spatial variables. For many systems, the Hamilton’s characteristic<br />

function (q P) separates <strong>in</strong>to a simple sum of terms each of which is a function of a s<strong>in</strong>gle variable. That<br />

is,<br />

(q α) = 1 ( 1 )+ 2 ( 2 )+····· ( ) (15.107)<br />

where each function <strong>in</strong> the summation on the right depends only on a s<strong>in</strong>gle variable. Then equation (15100)<br />

reduces to<br />

( 1 ; )= (15.108)<br />

1 <br />

where is the constant denot<strong>in</strong>g the total energy.<br />

Hamilton’s characteristic function (q P) can be used with equations (15101), (15102) (1591),<br />

(1592), and(1593) to derive<br />

=<br />

˙ = H =0<br />

<br />

H = + <br />

<br />

(q α)<br />

(q α)<br />

= (15.109)<br />

<br />

˙<br />

= H =0 (15.110)<br />

<br />

= − =0 (15.111)<br />

which has reduced the problem to a simple sum of one-dimensional first-order differential equations.<br />

If the variable is cyclic, then the Hamiltonian is not a function of and the term <strong>in</strong> Hamilton’s<br />

characteristic function equals = which separates out from the summation <strong>in</strong> equation 15107 That<br />

is, all cyclic variables can be factored out of (q α) which greatly simplifies solution of the Hamilton-Jacobi<br />

equation. As a consequence, the ability of the Hamilton-Jacobi method to make a canonical transformation to<br />

separate the system <strong>in</strong>to many cyclic or <strong>in</strong>dependent variables, which can be solved trivially, is a remarkably<br />

powerful way for solv<strong>in</strong>g the equations of motion <strong>in</strong> Hamiltonian mechanics.<br />

15.12 Example: Free particle<br />

Consider the motion of a free particle of mass <strong>in</strong> a force-free region. Then equation 1593 reduces to<br />

( 1 ; ; )+ <br />

1 =0<br />

S<strong>in</strong>ce no forces act, and the momentum p = ∇, thus the Hamilton-Jacobi equation reduces to<br />

1<br />

2 ∇2 + <br />

=0<br />

The Hamiltonian is time <strong>in</strong>dependent, thus equation 15101 applies<br />

()<br />

(q)= (q α) − (α)<br />

S<strong>in</strong>ce the Hamiltonian does not explicitly depend on the coord<strong>in</strong>ates ( ) then the coord<strong>in</strong>ates are cyclic<br />

and separation of the variables, 15107, gives that the action<br />

= α · r − <br />

For equation to be a solution of equation requires that<br />

= 1<br />

2 α2<br />

Therefore<br />

= α · r − 1<br />

2 α2 <br />

()<br />

()<br />

()


418 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

S<strong>in</strong>ce<br />

˙Q = S<br />

α = r− α <br />

the equation of motion and the conjugate momentum are given by<br />

r = ˙Q + α <br />

p = ∇ = α<br />

Thus the Hamilton-Jacobi relation has given both the equation of motion and the l<strong>in</strong>ear momentum p.<br />

15.13 Example: Po<strong>in</strong>t particle <strong>in</strong> a uniform gravitational field<br />

The Hamiltonian is<br />

= 1<br />

2 (2 + 2 + 2 )+<br />

S<strong>in</strong>ce the system is conservative, then the Hamilton-Jacobi equation can be written <strong>in</strong> terms of Hamilton’s<br />

characteristic function <br />

" µ<br />

= 1 2 µ 2 µ # 2 <br />

+ + + <br />

2 <br />

Assum<strong>in</strong>g that the variables can be separated = ()+ ()+() leads to<br />

= () = <br />

<br />

()<br />

= = <br />

<br />

= ()<br />

<br />

Thus by <strong>in</strong>tegration the total equals<br />

=<br />

Z <br />

Therefore us<strong>in</strong>g (15106) gives<br />

0<br />

+<br />

= − 0 =<br />

Z <br />

=<br />

0<br />

+<br />

Z <br />

0<br />

q<br />

2( − ) − 2 − 2 <br />

Z <br />

Z <br />

= constant =( − 0 ) −<br />

0<br />

Z <br />

= constant =( − 0 ) −<br />

0<br />

³q<br />

´<br />

2( − ) − 2 − 2 <br />

0<br />

<br />

q<br />

2( − ) − 2 − 2 <br />

<br />

q<br />

2( − ) − 2 − 2 <br />

<br />

q<br />

2( − ) − 2 − 2 <br />

If 0 0 0 is the position of the particle at time = 0 then = =0,andfrom(15106)<br />

− 0 =<br />

− 0 =<br />

− 0 =<br />

³ <br />

´<br />

( − 0 )<br />

³<br />

<br />

<br />

´<br />

( − 0 )<br />

<br />

⎛q<br />

⎞<br />

2( − ) − 2 − 2 <br />

⎝<br />

⎠ ( − 0 ) − 1 <br />

2 ( − 0) 2<br />

This corresponds to a parabola as should be expected for this trivial example.


15.4. HAMILTON-JACOBI THEORY 419<br />

15.14 Example: One-dimensional harmonic oscillator<br />

As discussed <strong>in</strong> example 1511 the Hamiltonian for the one-dimensional harmonic oscillator can be written<br />

as<br />

= 1 ¡<br />

2 + 2 2 2¢ = <br />

2<br />

q<br />

<br />

assum<strong>in</strong>g it is conservative and where =<br />

<br />

Hamilton’s characteristic function can be used where<br />

( ) = ( ) − <br />

= <br />

<br />

Insert<strong>in</strong>g the generalized momentum <strong>in</strong>to the Hamiltonian gives<br />

à ∙<br />

!<br />

¸2<br />

1<br />

+ 2 2 2 = <br />

2 <br />

Integration of this equation gives<br />

That is<br />

Note that<br />

This can be <strong>in</strong>tegrated to give<br />

= √ Z<br />

2<br />

= √ Z<br />

2<br />

( )<br />

<br />

r<br />

<br />

r Z 2<br />

=<br />

<br />

r<br />

1 − 2 2<br />

2<br />

1 − 2 2<br />

2<br />

− <br />

<br />

q − <br />

1 − 2 2<br />

2<br />

= 1 r !<br />

<br />

Ã<br />

arcs<strong>in</strong> 2<br />

+ 0<br />

2<br />

That is<br />

r<br />

2<br />

=<br />

2 s<strong>in</strong> ( − 0)<br />

This is the familiar solution of the undamped harmonic oscillator.<br />

15.15 Example: The central force problem<br />

The problem of a particle acted upon by a central force occurs frequently <strong>in</strong> physics. Consider the mass <br />

acted upon by a time-<strong>in</strong>dependent central potential energy () The Hamiltonian is time <strong>in</strong>dependent and<br />

can be written <strong>in</strong> spherical coord<strong>in</strong>ates as<br />

= 1<br />

2<br />

µ<br />

2 + 1 2 2 +<br />

<br />

1<br />

2 s<strong>in</strong> 2 2 + () =<br />

The time-<strong>in</strong>dependent Hamilton-Jacobi equation is conservative, thus<br />

" µ 2<br />

1<br />

+ 1 µ 2 µ # 2 1 <br />

2 2 +<br />

2 s<strong>in</strong> 2 + () =<br />

<br />

Try a separable solution for Hamilton’s characteristic function of the form<br />

= ()+Θ()+Φ()


420 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

The Hamilton-Jacobi equation then becomes<br />

" µ<br />

1<br />

2 <br />

This can be rearranged <strong>in</strong>to the form<br />

( " µ<br />

2 2 s<strong>in</strong> 2 1<br />

<br />

2 <br />

2<br />

+ 1 µ 2 Θ<br />

2 +<br />

<br />

1<br />

2 s<strong>in</strong> 2 <br />

µ # 2 Φ<br />

+ () =<br />

<br />

2<br />

+ 1 µ #<br />

)<br />

2 Θ<br />

2 + ()+ = −<br />

<br />

µ 2 Φ<br />

<br />

The left-hand side is <strong>in</strong>dependent of whereas the right-hand side is <strong>in</strong>dependent of and Both sides<br />

must equal a constant which is set to equal − 2 ,thatis<br />

1<br />

2<br />

" µ<br />

<br />

2<br />

+ 1 µ # 2 Θ<br />

2 <br />

2 + ()+<br />

<br />

2 2 s<strong>in</strong> 2 = <br />

µ 2 Φ<br />

= 2 <br />

<br />

The equation <strong>in</strong> and can be rearranged <strong>in</strong> the form<br />

" µ " 2 µΘ #<br />

2<br />

2 2 1 <br />

+ () − #<br />

= − + 2 <br />

2 <br />

s<strong>in</strong> 2 <br />

The left-hand side is <strong>in</strong>dependent of and the right-hand side is <strong>in</strong>dependent of so both must equal a<br />

constant which is set to be − 2 µ 2<br />

1 <br />

+ ()+ 2<br />

2 <br />

2 2 = <br />

µ 2 Θ<br />

+ 2 <br />

s<strong>in</strong> 2 = 2<br />

The variables now are completely separated and, by rearrangement plus <strong>in</strong>tegration, one obta<strong>in</strong>s<br />

() = √ Z r<br />

2 − () −<br />

2<br />

2 2 <br />

Z r<br />

Θ() = 2 −<br />

2 <br />

s<strong>in</strong> 2 <br />

Φ() = <br />

Substitut<strong>in</strong>g these <strong>in</strong>to = ()+Θ()+Φ() gives<br />

= √ Z r<br />

2 − () −<br />

Z r<br />

2<br />

2 2 + 2 −<br />

2 <br />

s<strong>in</strong> 2 + <br />

Hamilton’s characteristic function is the generat<strong>in</strong>g function from coord<strong>in</strong>ates ( ) to new<br />

coord<strong>in</strong>ates, which are cyclic, and new momenta that are constant and taken to be the separation constants<br />

<br />

= <br />

= √ r<br />

2 − () −<br />

= <br />

= r<br />

2 −<br />

= <br />

= <br />

2 <br />

s<strong>in</strong> 2 <br />

2<br />

2 2


15.4. HAMILTON-JACOBI THEORY 421<br />

Similarly, us<strong>in</strong>g (15109) gives the new coord<strong>in</strong>ates <br />

+ = <br />

= r <br />

2<br />

Z<br />

= <br />

= √ Z<br />

2<br />

<br />

= Z<br />

=<br />

<br />

<br />

q<br />

2 −<br />

<br />

q<br />

− () −<br />

<br />

q<br />

− () −<br />

2 <br />

s<strong>in</strong> 2 <br />

2<br />

2 2<br />

2<br />

2 2<br />

µ −<br />

2 2 <br />

+ <br />

µ Z −<br />

2 2 +<br />

<br />

q<br />

2 −<br />

2 <br />

s<strong>in</strong> 2 <br />

These equations lead to the elliptical, parabolic, or hyperbolic orbits discussed <strong>in</strong> chapter 11.<br />

15.16 Example: L<strong>in</strong>early-damped, one-dimensional, harmonic oscillator<br />

A canonical treatment of the l<strong>in</strong>early-damped harmonic oscillator provides an example that comb<strong>in</strong>es use<br />

of non-standard Lagrangian and Hamiltonians, a canonical transformation to an autonomous system, and<br />

use of Hamilton-Jacobi theory to solve this transformed system. It shows that Hamilton-Jacobi theory can be<br />

used to determ<strong>in</strong>e directly the solutions for the l<strong>in</strong>early-damped harmonic oscillator.<br />

Non-standard Hamiltonian:<br />

In chapter 35 the equation of motion for the l<strong>in</strong>early-damped, one-dimensional, harmonic oscillator was<br />

given to be<br />

£¨ + Γ ˙ + <br />

2<br />

2<br />

0 ¤ =0<br />

()<br />

Example 103 showed that three non-standard Lagrangians give equation of motion when used with the<br />

standard Euler-Lagrange variational equations. One of these was the Bateman[Bat31] time-dependent Lagrangian<br />

2 ( ˙ ) = £ 2 Γ ˙ 2 − 2 0 2¤<br />

()<br />

This Lagrangian gave the generalized momentum to be<br />

= 2<br />

˙ = ˙Γ<br />

()<br />

which was used with equation 153 to derive the Hamiltonian<br />

−Γ 2<br />

2 ( ) = ˙ − 2 ( ˙ ) =<br />

2 + 1 2 2 0 2 Γ<br />

()<br />

Note that both the Lagrangian and Hamiltonian are explicitly time dependent and thus they are not<br />

conserved quantities. This is as expected for this dissipative system.<br />

Hamilton-Jacobi theory:<br />

The form of the non-autonomous Hamiltonian () suggests use of the generat<strong>in</strong>g function for a canonical<br />

transformation to an autonomous Hamiltonian, for which H is a constant of motion.<br />

Then the canonical transformation gives<br />

( ) = 2 ( ) = Γ<br />

2 = ()<br />

= <br />

= Γ 2 ()<br />

= <br />

<br />

= <br />

Γ<br />

2<br />

Insert this canonical transformation <strong>in</strong>to the above Hamiltonian leads to the transformed Hamiltonian that<br />

is autonomous.<br />

H( )= 2 ( )+ 2<br />

= 2<br />

2 + Γ 2 + 2 0<br />

2 2 ()


422 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

That is, the transformed Hamiltonian H( ) is not explicitly time dependent, and thus is conserved.<br />

Expressed <strong>in</strong> the orig<strong>in</strong>al canonical variables ( ), the transformed Hamiltonian H( )<br />

H( )= 2<br />

2 −Γ + Γ 2 + 2 0<br />

2 2 Γ<br />

is a constant of motion which was not readily apparent when us<strong>in</strong>g the orig<strong>in</strong>al Hamiltonian. This unexpected<br />

result illustrates the usefulness of canonical transformations for solv<strong>in</strong>g dissipative systems. The Hamilton-<br />

Jacobi theory now can be used to solve the equations of motion for the transformed variables ( ) plus the<br />

transformed Hamiltonian H( ). The derivative of the generat<strong>in</strong>g function<br />

<br />

= <br />

()<br />

Use equation ( ) to substitute for <strong>in</strong> the Hamiltonian H( ) (equation ()), then the Hamilton-<br />

Jacobi method gives<br />

1<br />

2<br />

µ 2 <br />

+ Γ 2 <br />

+ 2 0<br />

2 2 + <br />

=0<br />

This equation is separable as described <strong>in</strong> 15107 and thus let<br />

( ) = ( ) − <br />

where is a separation constant. Then<br />

" µ #<br />

2<br />

1 <br />

+ Γ <br />

2 + 2 0<br />

2 2 = ()<br />

To simplify the equations def<strong>in</strong>e the variable as<br />

≡ √ 0 <br />

()<br />

then equation ( ) can be written as<br />

µ 2 <br />

+ <br />

+ ¡ 2 − ¢ =0<br />

()<br />

where = Γ 0<br />

and = 2<br />

0<br />

. Assume <strong>in</strong>itial conditions (0) = 0 and ˙(0) = 0<br />

For this case the separation constant 0 therefore 0. Note that equation ( ) isasimple<br />

second-order algebraic relation, the solution of which is<br />

v uut " µ # 2<br />

<br />

<br />

= − 2 ± − 1 − <br />

2 2<br />

()<br />

The choice of the sign is irrelevant for this case and thus the positive sign is chosen. There are three possible<br />

cases for the solution depend<strong>in</strong>g on whether the square-root term is real, zero, or imag<strong>in</strong>ary.<br />

<br />

Case 1:<br />

2 1, that is, <br />

2 0<br />

1<br />

r h<br />

Def<strong>in</strong>e = 1 − ¡ ¢<br />

2<br />

i<br />

2<br />

Then equation () canbe<strong>in</strong>tegratedtogive<br />

and<br />

This <strong>in</strong>tegral gives<br />

= − − 2<br />

4 + Z p(<br />

− 2 2 ) ()<br />

= <br />

= − + 1 0<br />

Z<br />

<br />

p<br />

( − 2 2 )<br />

s<strong>in</strong> −1 µ <br />

√<br />

<br />

<br />

= 0 ( + ) ≡ +


15.4. HAMILTON-JACOBI THEORY 423<br />

where<br />

µ <br />

s<br />

2 µ 2 Γ<br />

Γ<br />

= 0 = 0<br />

s1 − = 2 0<br />

2 − ()<br />

0 2<br />

Transform<strong>in</strong>g back to the orig<strong>in</strong>al variable gives<br />

() = − Γ<br />

2 s<strong>in</strong> ( + ) ()<br />

where and are given by the <strong>in</strong>itial conditions. Equation is identical to the solution for the underdamped<br />

l<strong>in</strong>early-damped l<strong>in</strong>ear oscillator given previously <strong>in</strong> equation 335.<br />

<br />

Case 2:<br />

2 =1, that is, Γ<br />

2 0<br />

=1<br />

r h<br />

In this case = 1 − ¡ ¢<br />

2<br />

i<br />

2<br />

=0and thus equation simplifies to<br />

= − − 2<br />

4 + √ <br />

and<br />

= <br />

= − + √<br />

0 <br />

Therefore the solution is<br />

() = − Γ<br />

2 ( + ) ()<br />

where F and G are constants given by the <strong>in</strong>itial conditions. This is the solution for the critically-damped<br />

l<strong>in</strong>early-damped, l<strong>in</strong>ear oscillator given previously <strong>in</strong> equation 338.<br />

<br />

Case 3:<br />

2 1, that is, Γ<br />

2 0<br />

1<br />

r h¡<br />

Def<strong>in</strong>e a real constant where = <br />

¢ 2<br />

i<br />

2 − 1 = , then<br />

= − − 2<br />

4 + Z p(<br />

+ 2 2 )<br />

Then<br />

This last <strong>in</strong>tegral gives<br />

= <br />

= − + 1 0<br />

Z<br />

<br />

p<br />

( + 2 2 )<br />

s<strong>in</strong>h −1 µ <br />

√<br />

<br />

<br />

= 0 ( + ) ≡ + <br />

where<br />

Then the orig<strong>in</strong>al variable gives<br />

s µ 2 <br />

= 0 = 0 − 1<br />

2 0<br />

() = − Γ<br />

2 s<strong>in</strong>h ( + ) ()<br />

This is the classic solution of the overdamped l<strong>in</strong>early-damped, l<strong>in</strong>ear harmonic oscillator given previously <strong>in</strong><br />

equation 337 The canonical transformation from a non-autonomous to an autonomous system allowed use<br />

of Hamiltonian mechanics to solve the damped oscillator problem.<br />

Note that this example used Bateman’s non-standard Lagrangian, and correspond<strong>in</strong>g Hamiltonian, for<br />

handl<strong>in</strong>g a dissipative l<strong>in</strong>ear oscillator system where the dissipation depends l<strong>in</strong>early on velocity. This nonstandard<br />

Lagrangian led to the correct equations of motion and solutions when applied us<strong>in</strong>g either the<br />

time-dependent Lagrangian, or time-dependent Hamiltonian, and these solutions agree with those given <strong>in</strong><br />

chapter 35 which were derived us<strong>in</strong>g Newtonian mechanics.


424 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

15.4.4 Visual representation of the action function .<br />

The important role of the action <strong>in</strong>tegral can be illum<strong>in</strong>ated<br />

by consider<strong>in</strong>g the case of a s<strong>in</strong>gle po<strong>in</strong>t mass<br />

mov<strong>in</strong>g <strong>in</strong> a time <strong>in</strong>dependent potential (). Then<br />

the action reduces to<br />

( ) = ( ) − (15.112)<br />

Let 1 = , 2 = 3 = 1 = 2 = 3 = .<br />

The momentum components are given by<br />

which corresponds to<br />

=<br />

( )<br />

<br />

(15.113)<br />

p = ∇ = ∇ (15.114)<br />

That is, the time-<strong>in</strong>dependent Hamilton-Jacobi equation<br />

is<br />

1<br />

2 |∇ |2 + () = (15.115)<br />

This implies that the particle momentum is given by<br />

the gradient of Hamilton’s characteristic function and is<br />

perpendicular to surfaces of constant as illustrated <strong>in</strong><br />

Figure 15.2: Surfaces of constant action <strong>in</strong>tegral S<br />

(dashed l<strong>in</strong>es) and the correspond<strong>in</strong>g particle momenta<br />

(solid l<strong>in</strong>es) with arrows show<strong>in</strong>g the direction.<br />

figure 152. The constant surfaces are time dependent as given by equation (15101) Thus, if at time<br />

=0the equi-action surface 0 ( ) = 0 ( )=0 then at =1the same surface 0 ( ) =0now<br />

co<strong>in</strong>cides with the 0 ( ) = surface etc. That is, the equi-action surfaces move through space separately<br />

from the motion of the s<strong>in</strong>gle po<strong>in</strong>t mass.<br />

The above pictorial representation is analogous to the situation for motion of a wavefront for electromagnetic<br />

waves <strong>in</strong> optics, or matter waves <strong>in</strong> quantum physics where the wave equation separates <strong>in</strong>to the form<br />

= 0 = 0 (k·r−) . Hamilton’s goal was to create a unified theory for optics that was equally applicable<br />

to particle motion <strong>in</strong> classical mechanics. Thus the optical-mechanical analogy of the Hamilton-Jacobi<br />

theory has culm<strong>in</strong>ated <strong>in</strong> a universal theory that describes wave-particle duality; this was a Holy Grail of<br />

classical mechanics s<strong>in</strong>ce Newton’s time. It played an important role <strong>in</strong> development of the Schröd<strong>in</strong>ger<br />

representation of quantum mechanics.<br />

15.4.5 Advantages of Hamilton-Jacobi theory<br />

Initially, only a few scientists, like Jacobi, recognized the advantages of Hamiltonian mechanics. In 1843<br />

Jacobi made some brilliant mathematical developments <strong>in</strong> Hamilton-Jacobi theory that greatly enhanced<br />

exploitation of Hamiltonian mechanics. Hamilton-Jacobi theory now serves as a foundation for contemporary<br />

physics, such as quantum and statistical mechanics. A major advantage of Hamilton-Jacobi theory, compared<br />

to other formulations of analytic mechanics, is that it provides a s<strong>in</strong>gle, first-order partial differential equation<br />

for the action which is a function of the generalized coord<strong>in</strong>ates q and time . The generalized momenta<br />

no longer appear explicitly <strong>in</strong> the Hamiltonian <strong>in</strong> equations 1594 1595. Note that the generalized momentum<br />

do not explicitly appear <strong>in</strong> the equivalent Euler-Lagrange equations of Lagrangian mechanics, but these<br />

comprise a system of second-order, partial differential equations for the time evolution of the generalized<br />

coord<strong>in</strong>ate q. Hamilton’s equations of motion are a system of 2 first-order equations for the time evolution<br />

of the generalized coord<strong>in</strong>ates and their conjugate momenta.<br />

An important advantage of the Hamilton-Jacobi theory is that it provides a formulation of classical<br />

mechanics <strong>in</strong> which motion of a particle can be represented by a wave. In this sense, the Hamilton-Jacobi<br />

equation fulfilled a long-held goal of theoretical physics, that dates back to Johann Bernoulli, of f<strong>in</strong>d<strong>in</strong>g an<br />

analogy between the propagation of light and the motion of a particle. This goal motivated Hamilton to<br />

develop Hamiltonian mechanics. A consequence of this wave-particle analogy is that the Hamilton-Jacobi<br />

formalism featured prom<strong>in</strong>ently <strong>in</strong> the derivation of the Schröd<strong>in</strong>ger equation dur<strong>in</strong>g the development of<br />

quantum-wave mechanics.


15.5. ACTION-ANGLE VARIABLES 425<br />

15.5 Action-angle variables<br />

15.5.1 Canonical transformation<br />

Systems possess<strong>in</strong>g periodic solutions are a ubiquitous feature <strong>in</strong> physics. The periodic motion can be either<br />

an oscillation, for which the trajectory <strong>in</strong> phase space is a closed loop (libration), or roll<strong>in</strong>g (rotational)<br />

motion as discussed <strong>in</strong> chapter 344. For many problems <strong>in</strong>volv<strong>in</strong>g periodic motion, the <strong>in</strong>terest often lies <strong>in</strong><br />

the frequencies of motion rather than the detailed shape of the trajectories <strong>in</strong> phase space. The action-angle<br />

variable approach uses a canonical transformation to action and angle variables which provide a powerful, and<br />

elegant method to exploit Hamiltonian mechanics. In particular, it can determ<strong>in</strong>e the frequencies of periodic<br />

motion without hav<strong>in</strong>g to calculate the exact trajectories for the motion. This method was <strong>in</strong>troduced by<br />

the French astronomer Ch. E. Delaunay(1816 − 1872) for applications to orbits <strong>in</strong> celestial mechanics, but<br />

it has equally important applications beyond celestial mechanics such as to bound solutions of the atom <strong>in</strong><br />

quantum mechanics.<br />

The action-angle method replaces the momenta <strong>in</strong> the Hamilton-Jacobi procedure by the action phase<br />

<strong>in</strong>tegral for the closed loop (libration) trajectory <strong>in</strong> phase space def<strong>in</strong>ed by<br />

I<br />

≡ (15.116)<br />

where for each cyclic variable the <strong>in</strong>tegral is taken over one complete period of oscillation. The cyclic variable<br />

is called the action variable where<br />

≡ 1<br />

2 = 1 I<br />

(15.117)<br />

2<br />

The canonical variable to the action variable I is the angle variable φ. Note that the name “action variable”<br />

is used to differentiate I from the action functional = R which has the same units; i.e. angular<br />

momentum.<br />

The general pr<strong>in</strong>ciple underly<strong>in</strong>g the use of action-angle variables is illustrated by consider<strong>in</strong>g one body,<br />

of mass , subject to a one-dimensional bound conservative potential energy (). The Hamiltonian is<br />

given by<br />

( ) = 2<br />

+ () (15.118)<br />

2<br />

This bound system has a ( ) phase space contour for each energy = <br />

( ) =± p 2( − ()) (15.119)<br />

For an oscillatory I system the two-valued momentum of equation 15119 is non-trivial to handle. By contrast,<br />

the area ≡ of the closed loop <strong>in</strong> phase space is a s<strong>in</strong>gle-valued scalar quantity that depends on <br />

I<br />

and (). Moreover, Liouville’s theorem states that the area of the closed contour <strong>in</strong> phase space ≡ <br />

is <strong>in</strong>variant to canonical transformations. These facts suggest the use of a new pair of conjugate variables,<br />

( ) where () uniquely labels the trajectory, and correspond<strong>in</strong>g area, of a closed loop <strong>in</strong> phase space<br />

for each value of , and the s<strong>in</strong>gle-valued function is a correspond<strong>in</strong>g angle that specifies the exact po<strong>in</strong>t<br />

along the phase-space contour as illustrated <strong>in</strong> Fig 153.<br />

For simplicity consider the l<strong>in</strong>ear harmonic oscillator where<br />

Then the Hamiltonian, 15118 equals<br />

() = 1 2 2 2 (15.120)<br />

( ) = 2<br />

2 + 1 2 2 2 (15.121)<br />

Hamilton’s equations of motion give that<br />

˙ = − <br />

= −2 (15.122)<br />

˙ = <br />

= (15.123)


426 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

The solution of equations 15122 and 15123 is of the form<br />

= cos(( − 0 )) (15.124)<br />

= − s<strong>in</strong> ( − 0 ) (15.125)<br />

where and 0 are <strong>in</strong>tegration constants. For the harmonic oscillator,<br />

equations 15124 and 15125 correspond to the usual elliptical contours<br />

<strong>in</strong> phase space, as illustrated <strong>in</strong> figure 153.<br />

The action-angle canonical transformation <strong>in</strong>volves mak<strong>in</strong>g the<br />

transform<br />

( ) → ( ) (15.126)<br />

where is def<strong>in</strong>ed by equation 15117 and the angle be<strong>in</strong>g the correspond<strong>in</strong>g<br />

canonical angle. The logical approach to this canonical<br />

transformation for the harmonic oscillator is to def<strong>in</strong>e and <strong>in</strong><br />

terms of and <br />

r<br />

2<br />

= cos (15.127)<br />

<br />

= √ 2 s<strong>in</strong> (15.128)<br />

Note that the Poisson bracket is unity<br />

[] ()<br />

=1<br />

which implies that the above transformation I<br />

is canonical, and thus<br />

the phase space area () ≡ 1<br />

2<br />

is conserved.<br />

For this canonical transformation the transformed Hamiltonian<br />

H ( ) is<br />

H ( ) = 1<br />

2 (2)s<strong>in</strong>2 + 1 2<br />

2<br />

2 cos2 = (15.129)<br />

Note that this Hamiltonian is a constant that is <strong>in</strong>dependent of the<br />

angle and thus Hamilton’s equations of motion give<br />

˙<br />

H ( )<br />

= − =0 (15.130)<br />

<br />

H ( )<br />

˙ = = (15.131)<br />

<br />

Thus we have mapped the harmonic oscillator to new coord<strong>in</strong>ates<br />

( ) where<br />

=<br />

H ( )<br />

= <br />

(15.132)<br />

= ( − 0 ) (15.133)<br />

Figure 15.3: The potential energy<br />

(), (upper) and correspond<strong>in</strong>g<br />

phase space ( ) (middle) for the<br />

harmonic oscillator at four equally<br />

spaced total energies . The correspond<strong>in</strong>g<br />

action-angles ( ) result<strong>in</strong>g<br />

from a canonical transformation<br />

of this system are shown <strong>in</strong> the lower<br />

plot.<br />

That is, the phase space has been mapped from ellipses, with area proportional to <strong>in</strong> the ( ) phase<br />

space, to a cyl<strong>in</strong>drical ( ) phase space where = <br />

are constant values that are <strong>in</strong>dependent of the angle,<br />

while <strong>in</strong>creases l<strong>in</strong>early with time. Thus the variables ( ) are periodic with modulus ∆ =2.<br />

( +2 ) = ( ) (15.134)<br />

( +2 ) = ( ) (15.135)<br />

The period of the periodic oscillatory motion is given simply by ∆ =2 = which is the well known result<br />

for the harmonic oscillator. Note that the action-angle variable canonical transformation has determ<strong>in</strong>ed<br />

the frequency of the periodic motion without solv<strong>in</strong>g the detailed trajectory of the motion.


15.5. ACTION-ANGLE VARIABLES 427<br />

The above example of the harmonic oscillator has shown that, for <strong>in</strong>tegrable periodic systems, it is<br />

possible to identify a canonical transformation to ( ) such that the Hamiltonian is <strong>in</strong>dependent of the<br />

angle which specifies the <strong>in</strong>stantaneous location on the constant energy contour . If the phase space<br />

contour is a separatrix, then it divides phase space <strong>in</strong>to <strong>in</strong>variant regions conta<strong>in</strong><strong>in</strong>g phase-space contours<br />

with differ<strong>in</strong>g behavior. The action-angle variables are not useful for separatrix contours. For roll<strong>in</strong>g motion,<br />

the system rotates with cont<strong>in</strong>uously <strong>in</strong>creas<strong>in</strong>g, or decreas<strong>in</strong>g angle, and there is no natural boundary for the<br />

action angle variable s<strong>in</strong>ce the phase space trajectory is cont<strong>in</strong>uous and not closed. However, the action-angle<br />

approach still is valid if the motion <strong>in</strong>volves periodic as well as roll<strong>in</strong>g motion.<br />

The example of the one-dimensional, one-body, harmonic oscillator can be expanded to the more general<br />

case for many bodies <strong>in</strong> three dimensions. This is illustrated by consider<strong>in</strong>g multiple periodic systems for<br />

which the Hamiltonian is conservative and where the equations of the canonical transformation are separable.<br />

The generalized momenta then can be written as<br />

= ( ; 1 2 )<br />

(15.136)<br />

<br />

for which each is a function of and the <strong>in</strong>tegration constants <br />

= ( 1 2 ) (15.137)<br />

The momentum ( 1 2 ) represents the trajectory of the system <strong>in</strong> the ( ) phase space that is<br />

characterized by Hamilton’s characteristic function ( ) Comb<strong>in</strong><strong>in</strong>g equations 15116 15136 gives<br />

I ( ; 1 2 )<br />

≡<br />

(15.138)<br />

<br />

S<strong>in</strong>ce is merely a variable of <strong>in</strong>tegration, each active action variable is a function of the constants<br />

of <strong>in</strong>tegration <strong>in</strong> the Hamilton-Jacobi equation. Because of the <strong>in</strong>dependence of the separable-variable pairs<br />

( ),the form <strong>in</strong>dependent functions of the and hence are suitable for use as a new set of constant<br />

momenta. Thus the characteristic function can be written as<br />

( 1 ; 1 )= X <br />

( ; 1 ) (15.139)<br />

while the Hamiltonian is only a function of the momenta ( 1 )<br />

The generalized coord<strong>in</strong>ate, conjugate to is known as the angle variable which is def<strong>in</strong>ed by the<br />

transformation equation<br />

= <br />

<br />

=<br />

X<br />

=1<br />

The correspond<strong>in</strong>g equation of motion for is given by<br />

( ; 1 )<br />

<br />

(15.140)<br />

˙ = () =2 ( 1 ) (15.141)<br />

<br />

where () are constant functions of the action variables with a solution<br />

=2 + (15.142)<br />

that is, they are l<strong>in</strong>ear functions of time The constants can be identified with the frequencies of the<br />

multiple periodic motions.<br />

The action-angle variables appear to be no different than a particular set of transformed coord<strong>in</strong>ates.<br />

Their merit appears when the physical <strong>in</strong>terpretation is assigned to . Consider the change as the <br />

are changed <strong>in</strong>f<strong>in</strong>itesimally<br />

= X <br />

= X 2 <br />

(15.143)<br />

<br />

<br />

<br />

The derivative with respect to vanishes except for the component of .Thusequation15143 reduces<br />

to


428 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

=<br />

X<br />

( ) (15.144)<br />

<br />

<br />

Therefore, the total change <strong>in</strong> as the system goes through one complete cycle is<br />

∆ = X I<br />

<br />

( ) =2 (15.145)<br />

<br />

<br />

<br />

where<br />

<br />

is outside the <strong>in</strong>tegral s<strong>in</strong>ce the are constants for cyclic motion. Thus ∆ =2 = where<br />

is the period for one cycle of oscillation, where the angular frequency is given by<br />

<br />

2 = = 1 <br />

(15.146)<br />

Thus the frequency associated with the periodic motion is the reciprocal of the period The secret here is<br />

that the derivative of with respect to the action variable given by equation (15141) directly determ<strong>in</strong>es<br />

thefrequencyoftheperiodicmotionwithouttheneedto solve the complete equations of motion. Note that<br />

multiple periodic motion can be represented by a Fourier expansion of the form<br />

∞X ∞X ∞X<br />

=<br />

1 <br />

2( 1 1 + 2 2 + 3 3 ++ )<br />

1=−∞ 2=−∞ =−∞<br />

(15.147)<br />

Although the action-angle approach to Hamilton-Jacobi theory does not produce complete equations of<br />

motion, it does provide the frequency decomposition that often is the physics of <strong>in</strong>terest. The reason that<br />

the powerful action-angle variable approach has been <strong>in</strong>troduced here is that it is used extensively <strong>in</strong> celestial<br />

mechanics. The action-angle concept also played a key role <strong>in</strong> the development of quantum mechanics, <strong>in</strong><br />

that Sommerfeld recognized that Bohr’s ad hoc assumption that angular momentum is quantized, could be<br />

expressed <strong>in</strong> terms of quantization of the angle variable as is mentioned <strong>in</strong> chapter 18.<br />

15.5.2 Adiabatic <strong>in</strong>variance of the action variables<br />

When the Hamiltonian depends on time it can be quite difficult to solve for the motion because it is difficult<br />

to f<strong>in</strong>d constants of motion for time-dependent systems. However, if the time dependence is sufficiently<br />

slow, that is, if the motion is adiabatic, then there exist dynamical variables that are almost constant which<br />

can be used to solve for the motion. In particular, such approximate constants are the familiar action-angle<br />

<strong>in</strong>tegrals. The adiabatic <strong>in</strong>variance of the action variables played an important role <strong>in</strong> the development of<br />

quantum mechanics dur<strong>in</strong>g the 1911 Solvay Conference. This was a time when physicists were grappl<strong>in</strong>g with<br />

the concepts of quantum mechanics. E<strong>in</strong>ste<strong>in</strong> used the follow<strong>in</strong>g classical mechanics example of adiabatic<br />

<strong>in</strong>variance, applied to the simple pendulum, <strong>in</strong> order to illustrate the concept of adiabatic <strong>in</strong>variance of the<br />

action. This example demonstrates the power of us<strong>in</strong>g action-angle variables.<br />

15.17 Example:Adiabatic<strong>in</strong>varianceforthesimplependulum<br />

Consider that the pendulum is made up of a po<strong>in</strong>t mass suspended from a pivot by a light str<strong>in</strong>g of<br />

length that is sw<strong>in</strong>g<strong>in</strong>g freely <strong>in</strong> a vertical plane. Derive the dependence of the amplitude of the oscillations<br />

, assum<strong>in</strong>g is small, if the str<strong>in</strong>g is very slowly shortened by a factor of 2, that is, assume that the change<br />

<strong>in</strong> length dur<strong>in</strong>g one period of the oscillation is very small.<br />

The tension <strong>in</strong> the str<strong>in</strong>g is given by<br />

= hcos i +<br />

+ * 2 ˙2<br />

<br />

Let the pendulum angle be oscillatory<br />

= 0 cos( + 0 )


15.5. ACTION-ANGLE VARIABLES 429<br />

Then the average mean square amplitude and velocity over one period are<br />

­<br />

<br />

2 ® = ­ [ 0 cos( + 0 )] 2® = 2 0<br />

2<br />

D˙2 E = ­ [− 0 s<strong>in</strong>( + 0 )] 2® = 2 2 0<br />

2<br />

S<strong>in</strong>ce, for the simple pendulum, 2 = ,thenthetension<strong>in</strong>thestr<strong>in</strong>g<br />

­<br />

<br />

2 ®<br />

= (1 −<br />

D˙2 E 2 )+ = (1 + 2 0<br />

4 )<br />

Assum<strong>in</strong>g that 0 is a small angle, and that the change <strong>in</strong> length −∆ is very small dur<strong>in</strong>g one period<br />

then the work done is<br />

∆ = ∆ = −∆ − 2 0<br />

∆ (a)<br />

4<br />

while the change <strong>in</strong> <strong>in</strong>ternal oscillator energy is<br />

∙<br />

)¸<br />

∆(−cos 0 )=∆ −(1 − 2 0<br />

= −∆ + 1 2 2 ∆(2 0)=−∆ + 1 2 2 0∆ + 0 ∆ 0<br />

The work done must balance the <strong>in</strong>crement <strong>in</strong> <strong>in</strong>ternal energy therefore<br />

or<br />

Therefore it follows that<br />

or<br />

0 ∆ 0 + 32 0∆<br />

4<br />

=0<br />

2 0∆ ln( 0 3 4 )=0<br />

( 0 3 4 )=constant (c)<br />

0 ∝ − 3 4<br />

Thus shorten<strong>in</strong>g the length of the pendulum str<strong>in</strong>g from to 2<br />

<strong>in</strong>creas<strong>in</strong>g by a factor 168.<br />

Consider the action-angle <strong>in</strong>tegral for one closed period = 2 <br />

I<br />

= <br />

I<br />

= 2 ˙ · ˙<br />

= 2 D˙2 E 2<br />

<br />

= 2 2 0<br />

= 1 2 <br />

2<br />

0 3 2 = constant<br />

(b)<br />

adiabatically corresponds to the amplitude<br />

for this problem<br />

where that last step is due to equation ().<br />

The above example shows that the action <strong>in</strong>tegral = , that is, it is <strong>in</strong>variant to an adiabatic<br />

change. In retrospect this result is as expected <strong>in</strong> that the action <strong>in</strong>tegral should be m<strong>in</strong>imized.


430 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

15.6 Canonical perturbation theory<br />

Most examples <strong>in</strong> classical mechanics discussed so far have been capable of exact solutions. In real life, the<br />

majority of problems cannot be solved exactly. For example, <strong>in</strong> celestial mechanics the two-body Kepler<br />

problem can be solved exactly, but solution of the three-body problem is <strong>in</strong>tractable. Typical systems <strong>in</strong><br />

celestial mechanics are never as simple as the two-body Kepler system because of the <strong>in</strong>fluence of additional<br />

bodies. Fortunately <strong>in</strong> most cases the <strong>in</strong>fluence of additional bodies is sufficiently small to allow use of<br />

perturbation theory. That is, the restricted three-body approximation can be employed for which the system<br />

is reduced to consider<strong>in</strong>g it as an exactly solvable two-body problem, subject to a small perturbation to this<br />

solvable two-body system. Note that even though the change <strong>in</strong> the Hamiltonian due to the perturb<strong>in</strong>g term<br />

may be small, the impact on the motion can be especially large near a resonance.<br />

Consider the Hamiltonian, subject to a time-dependent perturbation, is written as<br />

( ) = 0 ( )+∆( )<br />

where 0 ( ) designates the unperturbed Hamiltonian and ∆( ) designates the perturb<strong>in</strong>g term.<br />

For the unperturbed system the Hamilton-Jacobi equation is given by<br />

H( )= 0 ( 1 ; ; )+ <br />

1 =0<br />

(1590)<br />

where ( ) is the generat<strong>in</strong>g function for the canonical transformation ( ) → ( ). The perturbed<br />

( ) rema<strong>in</strong>s a canonical transformation, but the transformed Hamiltonian H( ) 6= 0.Thatis,<br />

H( )= 0 + ∆( )+ = ∆( ) (15.148)<br />

<br />

The equations of motion satisfied by the transformed variables now are<br />

˙ = ∆<br />

(15.149)<br />

<br />

˙<br />

= ∆<br />

<br />

These equations rema<strong>in</strong> as difficult to solve as the full Hamiltonian. However, the perturbation technique<br />

assumes that ∆ is small, and that one can neglect the change of ( ) over the perturb<strong>in</strong>g <strong>in</strong>terval.<br />

Therefore, to a first approximation, the unperturbed values of ∆<br />

<br />

and ∆<br />

<br />

canbeused<strong>in</strong>equations15149.<br />

A detailed explanation of canonical perturbation theory is presented <strong>in</strong> chapter 12 of Goldste<strong>in</strong>[Go50].<br />

15.18 Example: Harmonic oscillator perturbation<br />

(a) Consider first the Hamilton-Jacobi equation for the generat<strong>in</strong>g function ( ) for the case of a<br />

s<strong>in</strong>gle free particle subject to the Hamiltonian = 1 2 2 . F<strong>in</strong>d the canonical transformation = () and<br />

= () where and are the transformed coord<strong>in</strong>ate and momentum respectively.<br />

The Hamilton-Jacobi equation<br />

<br />

+ ( ) =0<br />

<br />

Us<strong>in</strong>g = <br />

<br />

<strong>in</strong> the Hamiltonian = 1 2 2 gives<br />

<br />

+ 1 2<br />

µ 2 <br />

=0<br />

<br />

S<strong>in</strong>ce does not depend on explicitly, then the two terms on the left hand side of the equation can be<br />

set equal to − respectively, where is at most a function of . Then the generat<strong>in</strong>g function is<br />

= p 2 − <br />

Set = √ 2 then the generat<strong>in</strong>g function can be written as<br />

= − 1 2 2


15.6. CANONICAL PERTURBATION THEORY 431<br />

The constant can be identified with the new momentum Then the transformation equations become<br />

= <br />

= <br />

= <br />

= <br />

= − = <br />

That is<br />

= + <br />

which corresponds to motion with a uniform velocity <strong>in</strong> the system.<br />

(b) Consider that the Hamiltonian is perturbed by addition of potential = 2<br />

2<br />

which corresponds to the<br />

harmonic oscillator. Then<br />

= 1 2 2 + 2<br />

2<br />

Consider the transformed Hamiltonian<br />

H = + <br />

= 1 2 2 + 2<br />

2 − 2<br />

2 = 2<br />

2 = 1 ( + )2<br />

2<br />

Hamilton’s equations of motion<br />

give that<br />

˙ = H<br />

<br />

˙ = ( + ) <br />

˙ = − ( + )<br />

˙ = − H<br />

<br />

These two equations can be solved to give<br />

¨ + =0<br />

which is the equation of a harmonic oscillator show<strong>in</strong>g that is harmonic of the form = 0 s<strong>in</strong> ( + )<br />

where 0 are constants of motion. Thus<br />

The transformation equations then give<br />

= − ˙ − = − 0 [cos( + )+ s<strong>in</strong>( + )]<br />

= = 0 s<strong>in</strong> ( + )<br />

= + = − ˙ = − 0 cos( + )<br />

Hence the solution for the perturbed system is harmonic, which is to be expected s<strong>in</strong>ce the potential has a<br />

quadratic dependence of position.<br />

15.19 Example: L<strong>in</strong>dblad resonance <strong>in</strong> planetary and galactic motion<br />

Use of canonical perturbation theory <strong>in</strong> celestial mechanics has been exploited by Professor Alice Quillen<br />

and her group. They comb<strong>in</strong>e use of action-angle variables and Hamilton-Jacobi theory to <strong>in</strong>vestigate the role<br />

of L<strong>in</strong>dblad resonance to planetary motion, and also for stellar motion <strong>in</strong> galaxies. A L<strong>in</strong>dblad resonance<br />

is an orbital resonance <strong>in</strong> which the orbital period of a celestial body is a simple multiple of some forc<strong>in</strong>g<br />

frequency. Even for very weak perturb<strong>in</strong>g forces, such resonance behavior can lead to orbit capture and chaotic<br />

motion.<br />

For planetary motion the planet masses are about 11000 that of the central star, so the perturbations<br />

to Kepler orbits are small. However, L<strong>in</strong>dblad resonance for planetary motion led to Saturn’s r<strong>in</strong>gs which<br />

result from perturbations produced by the moons of Saturn that skulpt and clear dust r<strong>in</strong>gs. Stellar orbits <strong>in</strong><br />

disk galaxies are perturbed a few percent by non axially-symmetric galactic features such as spiral arms or<br />

bars. L<strong>in</strong>dblad resonances perturb stellar motion and drive spiral density waves at distances from the center<br />

of a galactic disk where the natural frequency of the radial component of a star’s orbital velocity is close to<br />

the frequency of the fluctuations <strong>in</strong> the gravitational field due to passage through spiral arms or bars. If a<br />

stars orbital speed around a galactic center is greater than that of the part of a spiral arm through which it is<br />

travers<strong>in</strong>g, then an <strong>in</strong>ner L<strong>in</strong>dblad resonance occurs which speeds up the star’s orbital speed mov<strong>in</strong>g the orbit<br />

outwards. If the orbital speed is less than that of a spiral arm, an <strong>in</strong>ner L<strong>in</strong>dblad resonance occurs caus<strong>in</strong>g<br />

<strong>in</strong>ward movement of the orbit.


432 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

15.7 Symplectic representation<br />

The Hamilton’s first-order equations of motion are symmetric if the generalized and constra<strong>in</strong>t force terms,<br />

<strong>in</strong> equation 159 are excluded.<br />

˙q = <br />

p<br />

− ṗ = <br />

q<br />

This stimulated attempts to treat the canonical variables (q p) <strong>in</strong> a symmetric form us<strong>in</strong>g group theory.<br />

Some graduate textbooks <strong>in</strong> classical mechanics have adopted use of symplectic symmetry <strong>in</strong> order to unify<br />

the presentation of Hamiltonian mechanics. For a system of degrees of freedom, a column matrix η is<br />

constructed that has 2 elements where<br />

Therefore the column matrix<br />

µ <br />

η<br />

<br />

= + = ≤ (15.150)<br />

= <br />

<br />

µ <br />

= ≤ (15.151)<br />

η<br />

+<br />

<br />

ThesymplecticmatrixJ is def<strong>in</strong>ed as be<strong>in</strong>g a 2 by 2 skew-symmetric, orthogonal matrix that is broken<br />

<strong>in</strong>to four × null or unit matrices accord<strong>in</strong>g to the scheme<br />

µ <br />

[0] +[1]<br />

J =<br />

(15.152)<br />

− [1] [0]<br />

where [0] is the -dimension null matrix, for which all elements are zero. Also [1] is the -dimensional unit<br />

matrix, for which the diagonal matrix elements are unity and all off-diagonal matrix elements are zero. The<br />

J matrix accounts for the opposite signs used <strong>in</strong> the equations for ˙q and ṗ. The symplectic representation<br />

allows the Hamilton’s equations of motion to be written <strong>in</strong> the compact form<br />

˙η = J <br />

η<br />

(15.153)<br />

This textbook does not use the elegant symplectic representation s<strong>in</strong>ce it ignores the important generalized<br />

forces and Lagrange multiplier forces.<br />

15.8 Comparison of the Lagrangian and Hamiltonian formulations<br />

Common features<br />

The discussion of Lagrangian and Hamiltonian dynamics has illustrated the power of such algebraic formulations.<br />

Both approaches are based on application of variational pr<strong>in</strong>ciples to scalar energy which gives the<br />

freedom to concentrate solely on active forces and to ignore <strong>in</strong>ternal forces. Both methods can handle manybody<br />

systems and exploit canonical transformations, which are impractical or impossible us<strong>in</strong>g the vectorial<br />

Newtonian mechanics. These algebraic approaches simplify the calculation of the motion for constra<strong>in</strong>ed<br />

systems by represent<strong>in</strong>g the vector force fields, as well as the correspond<strong>in</strong>g equations of motion, <strong>in</strong> terms of<br />

either the Lagrangian function (q ˙q) or the action functional (q p) which are related by the def<strong>in</strong>ite<br />

<strong>in</strong>tegral<br />

Z 2<br />

(q p) = (q ˙q) (151)<br />

1<br />

The Lagrangian function (q ˙q) and the action functional (q p) are scalar functions under rotation,<br />

but they determ<strong>in</strong>e the vector force fields and the correspond<strong>in</strong>g equations of motion. Thus the use of<br />

rotationally-<strong>in</strong>variant functions (q ˙q) and (q p) provide a simple representation of the vector force<br />

fields. This is analogous to the use of scalar potential fields (q) to represent the electrostatic and gravitational<br />

vector force fields. Like scalar potential fields, Lagrangian and Hamiltonian mechanics represents the<br />

observables as derivatives of (q ˙q) and (q p) and the absolute values of (q ˙q) and (q p) are<br />

undef<strong>in</strong>ed; only differences <strong>in</strong> (q ˙q) and (q p) are observable. For example, the generalized momenta<br />

are given by the derivatives ≡ <br />

˙ <br />

and = <br />

<br />

. The physical significance of the least action (q α) is


15.8. COMPARISON OF THE LAGRANGIAN AND HAMILTONIAN FORMULATIONS 433<br />

illustrated when the canonically transformed momenta P = α is a constant. Then the generalized momenta<br />

and the Hamilton-Jacobi equation, imply that the total time derivative of the action equals<br />

<br />

= <br />

<br />

˙ + <br />

= − = (15.154)<br />

The <strong>in</strong>def<strong>in</strong>ite <strong>in</strong>tegral of this equation reproduces the def<strong>in</strong>ite <strong>in</strong>tegral (151) to with<strong>in</strong> an arbitrary constant,<br />

i.e.<br />

Z<br />

(q p) = (q ˙q) + constant (15.155)<br />

Lagrangian formulation:<br />

Consider a system with <strong>in</strong>dependent generalized coord<strong>in</strong>ates, plus constra<strong>in</strong>t forces that are not required<br />

to be known. The Lagrangian approach can reduce the system to a m<strong>in</strong>imal system of = − <strong>in</strong>dependent<br />

generalized coord<strong>in</strong>ates lead<strong>in</strong>g to = − second-order differential equations. By comparison,<br />

the Newtonian approach uses + unknowns. Alternatively, the Lagrange multipliers approach allows<br />

determ<strong>in</strong>ation of the holonomic constra<strong>in</strong>t forces result<strong>in</strong>g <strong>in</strong> = + second order equations to determ<strong>in</strong>e<br />

= + unknowns. The Lagrangian potential function is limited to conservative forces, but generalized<br />

forces can be used to handle non-conservative and non-holonomic forces. The advantage of the Lagrange<br />

equations of motion is that they can deal with any type of force, conservative or non-conservative, and<br />

they directly determ<strong>in</strong>e , ˙ rather than which then requires relat<strong>in</strong>g to ˙. The Lagrange approach is<br />

superior to the Hamiltonian approach if a numerical solution is required for typical undergraduate problems<br />

<strong>in</strong> classical mechanics. However, Hamiltonian mechanics has a clear advantage for address<strong>in</strong>g more profound<br />

and philosophical questions <strong>in</strong> physics.<br />

Hamiltonian formulation:<br />

For a system with <strong>in</strong>dependent generalized coord<strong>in</strong>ates, and constra<strong>in</strong>t forces, the Hamiltonian approach<br />

determ<strong>in</strong>es 2 first-order differential equations. In contrast to Lagrangian mechanics, where the Lagrangian<br />

is a function of the coord<strong>in</strong>ates and their velocities, the Hamiltonian uses the variables q and p, rather<br />

than velocity. The Hamiltonian has twice as many <strong>in</strong>dependent variables as the Lagrangian which is a great<br />

advantage, not a disadvantage, s<strong>in</strong>ce it broadens the realm of possible transformations that can be used to<br />

simplify the solutions. Hamiltonian mechanics uses the conjugate coord<strong>in</strong>ates q p correspond<strong>in</strong>g to phase<br />

space. This is an advantage <strong>in</strong> most branches of physics and eng<strong>in</strong>eer<strong>in</strong>g. Compared to Lagrangian mechanics,<br />

Hamiltonian mechanics has a significantly broader arsenal of powerful techniques that can be exploited to<br />

obta<strong>in</strong> an analytical solution of the <strong>in</strong>tegrals of the motion for complicated systems. These techniques<br />

<strong>in</strong>clude, the Poisson bracket formulation, canonical transformations, the Hamilton-Jacobi approach, the<br />

action-angle variables, and canonical perturbation theory. In addition, Hamiltonian dynamics also provides<br />

a means of determ<strong>in</strong><strong>in</strong>g the unknown variables for which the solution assumes a soluble form, and it is<br />

ideal for study of the fundamental underly<strong>in</strong>g physics <strong>in</strong> applications to other fields such as quantum or<br />

statistical physics. However, the Hamiltonian approach endemically assumes that the system is conservative<br />

putt<strong>in</strong>g it at a disadvantage with respect to the Lagrangian approach. The appeal<strong>in</strong>g symmetry of the<br />

Hamiltonian equations, plus their ability to utilize canonical transformations, makes it the formalism of<br />

choice for exam<strong>in</strong>ation of system dynamics. For example, Hamilton-Jacobi theory, action-angle variables<br />

and canonical perturbation theory are used extensively to solve complicated multibody orbit perturbations<br />

<strong>in</strong> celestial mechanics by f<strong>in</strong>d<strong>in</strong>g a canonical transformation that transforms the perturbed Hamiltonian to<br />

a solved unperturbed Hamiltonian.<br />

The Hamiltonian formalism features prom<strong>in</strong>ently <strong>in</strong> quantum mechanics s<strong>in</strong>ce there are well established<br />

rules for transform<strong>in</strong>g the classical coord<strong>in</strong>ates and momenta <strong>in</strong>to l<strong>in</strong>ear operators used <strong>in</strong> quantum mechanics.<br />

The variables q ˙q used <strong>in</strong> Lagrangian mechanics do not have simple analogs <strong>in</strong> quantum physics.<br />

As a consequence, the Poisson bracket formulation, and action-angle variables of Hamiltonian mechanics<br />

played a key role <strong>in</strong> development of matrix mechanics by Heisenberg, Born, and Dirac, while the Hamilton-<br />

Jacobi formulation played a key role <strong>in</strong> development of Schröd<strong>in</strong>ger’s wave mechanics. Similarly, Hamiltonian<br />

mechanics is the preem<strong>in</strong>ent variational approached used <strong>in</strong> statistical mechanics.


434 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

15.9 Summary<br />

This chapter has gone beyond what is normally covered <strong>in</strong> an undergraduate course <strong>in</strong> classical mechanics,<br />

<strong>in</strong> order to illustrate the power of the remarkable arsenal of methods available for solution of the equations of<br />

motion us<strong>in</strong>g Hamiltonian mechanics. This has <strong>in</strong>cluded the Poisson bracket representation of Hamiltonian<br />

formulation of mechanics, canonical transformations, Hamilton-Jacobi theory, action-angle variables, and<br />

canonical perturbation theory. The purpose was to illustrate the power of variational pr<strong>in</strong>ciples <strong>in</strong> Hamiltonian<br />

mechanics and how they relate to fields such as quantum mechanics. The follow<strong>in</strong>g are the key po<strong>in</strong>ts<br />

made <strong>in</strong> this chapter.<br />

Poisson brackets: The elegant and powerful Poisson bracket formalism of Hamiltonian mechanics was<br />

<strong>in</strong>troduced. The Poisson bracket of any two cont<strong>in</strong>uous functions of generalized coord<strong>in</strong>ates ( ) and<br />

( ) is def<strong>in</strong>ed to be<br />

[ ] <br />

≡ X µ <br />

− <br />

<br />

(1513)<br />

<br />

<br />

The fundamental Poisson brackets equal<br />

[ ]=0 (1521)<br />

[ ]=0 (1522)<br />

[ ]=− [ ]= (1523)<br />

The Poisson bracket is <strong>in</strong>variant to a canonical transformation from ( ) to ( ). Thatis<br />

[ ] <br />

= X <br />

µ <br />

− <br />

<br />

=[ ]<br />

<br />

<br />

(1532)<br />

There is a one-to-one correspondence between the commutator and Poisson Bracket of two <strong>in</strong>dependent<br />

functions,<br />

( 1 1 − 1 1 )= [ 1 1 ] (1538)<br />

where is an <strong>in</strong>dependent constant. In particular 1 1 commute of the Poisson Bracket [ 1 1 ]=0.<br />

Poisson Bracket representation of Hamiltonian mechanics: It has been shown that the Poisson<br />

bracket formalism conta<strong>in</strong>s the Hamiltonian equations of motion and is <strong>in</strong>variant to canonical transformations.<br />

Also this formalism extends Hamilton’s canonical equations to non-commut<strong>in</strong>g canonical variables.<br />

Hamilton’s equations of motion can be expressed directly <strong>in</strong> terms of the Poisson brackets<br />

˙ =[ ]= <br />

<br />

˙ =[ ]=− <br />

<br />

(1557)<br />

(1558)<br />

An important result is that the total time derivative of any operator is given by<br />

<br />

= +[ ]<br />

(1545)<br />

Poisson brackets provide a powerful means of determ<strong>in</strong><strong>in</strong>g which observables are time <strong>in</strong>dependent and<br />

whether different observables can be measured simultaneously with unlimited precision. It was shown that<br />

the Poisson bracket is <strong>in</strong>variant to canonical transformations, which is a valuable feature for Hamiltonian<br />

mechanics. Poisson brackets were used to prove Liouville’s theorem which plays an important role <strong>in</strong> the use<br />

of Hamiltonian phase space <strong>in</strong> statistical mechanics. The Poisson bracket is equally applicable to cont<strong>in</strong>uous<br />

solutions <strong>in</strong> classical mechanics as well as discrete solutions <strong>in</strong> quantized systems.


15.9. SUMMARY 435<br />

Canonical transformations: A transformation between a canonical set of variables ( ) with Hamiltonian<br />

( ) to another set of canonical variable ( ) with Hamiltonian H( ) can be achieved<br />

us<strong>in</strong>g a generat<strong>in</strong>g functions such that<br />

H( ) =( )+ <br />

<br />

(1589)<br />

Possible generat<strong>in</strong>g functions are summarized <strong>in</strong> the follow<strong>in</strong>g table.<br />

Generat<strong>in</strong>g function Generat<strong>in</strong>g function derivatives Trivial special case<br />

= 1 (q Q)<br />

= 1<br />

<br />

= − 1<br />

<br />

1 = = = − <br />

= 2 (q P) − Q · P = 2<br />

<br />

= 2<br />

<br />

2 = = = <br />

= 3 (p Q)+q · p = − 3<br />

<br />

= − 3<br />

<br />

3 = = − = − <br />

= 4 (p P)+q · p − Q · P = − 4<br />

<br />

= 4<br />

<br />

1 = = = − <br />

If the canonical transformation makes H( ) =0then the conjugate variables ( ) are constants<br />

of motion. Similarly if H( ) is a cyclic function then the correspond<strong>in</strong>g are constants of motion.<br />

Hamilton-Jacobi theory: Hamilton-Jacobi theory determ<strong>in</strong>es the generat<strong>in</strong>g function required to perform<br />

canonical transformations that leads to a powerful method for obta<strong>in</strong><strong>in</strong>g the equations of motion for<br />

a system. The Hamilton-Jacobi theory uses the action function ≡ 2 as a generat<strong>in</strong>g function, and the<br />

canonical momentum is given by<br />

= <br />

(154)<br />

<br />

This can be used to replace <strong>in</strong> the Hamiltonian lead<strong>in</strong>g to the Hamilton-Jacobi equation<br />

(; <br />

; )+ =0<br />

(1594)<br />

Solutions of the Hamilton-Jacobi equation were obta<strong>in</strong>ed by separation of variables. The close opticalmechanical<br />

analogy of the Hamilton-Jacobi theory is an important advantage of this formalism that led to<br />

it play<strong>in</strong>g a pivotal role <strong>in</strong> the development of wave mechanics by Schröd<strong>in</strong>ger.<br />

Action-angle variables:<br />

( ) where<br />

The action-angle variables exploits a canonical transformation from ( ) →<br />

≡ 1<br />

2 = 1 I<br />

<br />

(15117)<br />

2<br />

For periodic motion the phase-space trajectory is closed with area given by andthisareaisconservedfor<br />

the above canonical transformation. For a conserved Hamiltonian the action variable is <strong>in</strong>dependent of<br />

the angle variable . The time dependence of the angle variable directly determ<strong>in</strong>es the frequency of the<br />

periodic motion without recourse to calculation of the detailed trajectory of the periodic motion.<br />

Canonical perturbation theory: Canonical perturbation theory is a valuable method of handl<strong>in</strong>g multibody<br />

<strong>in</strong>teractions. The adiabatic <strong>in</strong>variance of the action-angle variables provides a powerful approach for<br />

exploit<strong>in</strong>g canonical perturbation theory.<br />

Comparison of Lagrangian and Hamiltonian formulations: Theremarkablepower,and<strong>in</strong>tellectual<br />

beauty, provided by use of variational pr<strong>in</strong>ciples to exploit the underly<strong>in</strong>g pr<strong>in</strong>ciples of natural economy <strong>in</strong><br />

nature, has had a long and rich history. It has led to profound developments <strong>in</strong> many branches of theoretical<br />

physics. However, it is noted that although the above algebraic formulations of classical mechanics have been<br />

used for over two centuries, the important limitations of these algebraic formulations to non-l<strong>in</strong>ear systems<br />

rema<strong>in</strong> a challenge that still is be<strong>in</strong>g addressed.<br />

It has been shown that the Lagrangian and Hamiltonian formulations represent the vector force fields,<br />

and the correspond<strong>in</strong>g equations of motion, <strong>in</strong> terms of the Lagrangian function (q ˙q) or the action


436 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

functional (q p) which are scalars under rotation. The Lagrangian function (q ˙q) is related to the<br />

action functional (q p) by<br />

Z 2<br />

(q p) = (q ˙q) (151)<br />

1<br />

These functions are analogous to electric potential, <strong>in</strong> that the observables are derived by tak<strong>in</strong>g derivatives<br />

of the Lagrangian function (q ˙q) or the action functional (q p). The Lagrangian formulation is more<br />

convenient for deriv<strong>in</strong>g the equations of motion for simple mechanical systems. The Hamiltonian formulation<br />

has a greater arsenal of techniques for solv<strong>in</strong>g complicated problems plus it uses the canonical variables ( )<br />

which are the variables of choice for applications to quantum mechanics and statistical mechanics.


15.9. SUMMARY 437<br />

Problems<br />

1. Poisson brackets are a powerful means of elucidat<strong>in</strong>g when observables are constant of motion and whether<br />

two observables can be simultaneously measured with unlimited precision. Consider a spherically symmetric<br />

Hamiltonian<br />

Ã<br />

= 1 2 + 2 <br />

2 2 +<br />

2 <br />

2 s<strong>in</strong> 2 <br />

!<br />

+ ()<br />

for a mass where ( is a central potential. Use the Poisson bracket plus the time dependence to determ<strong>in</strong>e<br />

the follow<strong>in</strong>g:<br />

(a) Does commute with and is it a constant of motion?<br />

(b) Does 2 +<br />

2 <br />

commute with and is it a constant of motion?<br />

s<strong>in</strong> 2 <br />

(c) Does commute with and is it a constant of motion?<br />

(d) Does commute with and what does the result imply?<br />

2. Consider the Poisson brackets for angular momentum L<br />

(a) Show { } = , where the Levi-Cevita tensor is,<br />

(b) Show { } = .<br />

⎧<br />

⎨ +1 if are cyclically permuted<br />

= −1 if are anti-cyclically permuted<br />

⎩<br />

0 if = or = or = <br />

(c) Show { } = . The follow<strong>in</strong>g identity may be useful: = − .<br />

(d) Show { 2 } =0.<br />

3. Consider the Hamiltonian of a two-dimensional harmonic oscillator,<br />

What condition is satisfied if 2 a conserved quantity?<br />

= p2<br />

2 + 1 2 ¡ 2 11 2 + 2 22<br />

2 ¢<br />

4. Consider the motion of a particle of mass <strong>in</strong> an isotropic harmonic oscillator potential = 1 2 2 and take<br />

theorbitalplanetobethe − plane. The Hamiltonian is then<br />

Introduce the three quantities<br />

with =<br />

≡ 0 = 1<br />

2 (2 + 2 )+ 1 2 (2 + 2 )<br />

1 = 1<br />

2 (2 − 2 )+ 1 2 (2 − 2 )<br />

2 = 1 + <br />

3 = ( − )<br />

q<br />

<br />

<br />

. Use Poisson brackets to solve the follow<strong>in</strong>g:<br />

(a) Show that [ 0 ]=0for =1 2 3 prov<strong>in</strong>g that ( 1 2 3 ) are constants of motion.


438 CHAPTER 15. ADVANCED HAMILTONIAN MECHANICS<br />

(b) Show that<br />

[ 1 2 ] = 2 3<br />

[ 2 3 ] = 2 1<br />

[ 3 1 ] = 2 2<br />

so that (2) −1 ( 1 2 3 ) have the same Poisson bracket relations as the components of a 3-dimensional<br />

angular momentum.<br />

2 0 = 2 1 + 2 2 + 2 3<br />

5. Assume that the transformation equations between the two sets of coord<strong>in</strong>ates ( ) and ( ) are<br />

= ln(1+ 1 2 cos )<br />

= 2(1+ 1 1<br />

2 cos ) 2 s<strong>in</strong> )<br />

(a) Assum<strong>in</strong>g that are canonical variables, i.e. [ ] =1, show directly from the above transformation<br />

equations that are canonical variables.<br />

(b) Show that the generat<strong>in</strong>g function that generates this transformation between the two sets of canonical<br />

variables is<br />

3 = −[ − 1] 2 tan <br />

6. Consider a bound two-body system compris<strong>in</strong>g a mass <strong>in</strong> an orbit at a distance from a mass . The<br />

attractive central force b<strong>in</strong>d<strong>in</strong>g the two-body system is<br />

F = 2ˆr<br />

where is negative. Use Poisson brackets to prove that the eccentricity vector = × + ˆ is a conserved<br />

quantity.<br />

7. Considerthecaseofas<strong>in</strong>glemass where the Hamiltonian = 1 2 2 .<br />

(a) Use the generat<strong>in</strong>g function () to solve the Hamilton-Jacobi equation with the canonical transformation<br />

= ( ) and = ( ) and determ<strong>in</strong>e the equations relat<strong>in</strong>g the ( ) variables to the<br />

transformed coord<strong>in</strong>ate and momentum ( ).<br />

(b) If there is a perturb<strong>in</strong>g Hamiltonian ∆ = 1 2 2 ,then will not be constant. Express the transformed<br />

Hamiltonian (us<strong>in</strong>g the transformation given above <strong>in</strong> terms of and ). Solvefor() and ()<br />

and show that the perturbed solution [()()] [()()] is simple harmonic.


Chapter 16<br />

Analytical formulations for cont<strong>in</strong>uous<br />

systems<br />

16.1 Introduction<br />

Lagrangian and Hamiltonian mechanics have been used to determ<strong>in</strong>e the equations of motion for discrete systems<br />

hav<strong>in</strong>g a f<strong>in</strong>ite number of discrete variables where 1 ≤ ≤ . There are important classes of systems<br />

where it is more convenient to treat the system as be<strong>in</strong>g cont<strong>in</strong>uous. For example, the <strong>in</strong>teratomic spac<strong>in</strong>g <strong>in</strong><br />

solids is a few 10 −10 which is negligible compared with the size of typical macroscopic, three-dimensional<br />

solid objects. As a consequence, for wavelengths much greater than the atomic spac<strong>in</strong>g <strong>in</strong> solids, it is useful<br />

to treat macroscopic crystall<strong>in</strong>e lattice systems as cont<strong>in</strong>uous three-dimensional uniform solids, rather<br />

than as three-dimensional discrete lattice cha<strong>in</strong>s. Fluid and gas dynamics are other examples of cont<strong>in</strong>uous<br />

mechanical systems. Another important class of cont<strong>in</strong>uous systems <strong>in</strong>volves the theory of fields, such as<br />

electromagnetic fields. Lagrangian and Hamiltonian mechanics of the cont<strong>in</strong>ua extend classical mechanics<br />

<strong>in</strong>to the advanced topic of field theory. This chapter goes beyond the scope of a typical undergraduate<br />

classical mechanics course <strong>in</strong> order to provide a brief glimpse of how Lagrangian and Hamiltonian mechanics<br />

can underlie advanced and important aspects of the mechanics of the cont<strong>in</strong>ua, <strong>in</strong>clud<strong>in</strong>g field theory.<br />

16.2 The cont<strong>in</strong>uous uniform l<strong>in</strong>ear cha<strong>in</strong><br />

The Lagrangian for the discrete lattice cha<strong>in</strong>, for longitud<strong>in</strong>al modes, is given by equation 1476 to be<br />

= 1 +1<br />

X ³ 2´<br />

˙ 2 − ( −1 − )<br />

2<br />

=1<br />

(16.1)<br />

where the masses are attached <strong>in</strong> series to +1 identical spr<strong>in</strong>gs of length and spr<strong>in</strong>g constant . Assume<br />

that the spr<strong>in</strong>g has a uniform cross-section area and length Then each spr<strong>in</strong>g volume element ∆ = <br />

has a mass , that is, the volume mass density = ∆<br />

or = ∆. Chapter 1653 will show that the<br />

spr<strong>in</strong>g constant = <br />

<br />

where is Young’s modulus, is the cross sectional area of the cha<strong>in</strong> element, and<br />

is the length of the element. Then the spr<strong>in</strong>g constant can be written as = ∆<br />

<br />

. Therefore equation<br />

2<br />

161 can be expressed as a sum over volume elements ∆ = <br />

Ã<br />

= 1 +1<br />

X<br />

µ ! 2<br />

˙ 2 −1 − <br />

− <br />

∆ (16.2)<br />

2<br />

<br />

=1<br />

In the limit that →∞and the spac<strong>in</strong>g = → 0 then the summation <strong>in</strong> equation 162 can be written<br />

as a volume <strong>in</strong>tegral where = is the distance along the l<strong>in</strong>ear cha<strong>in</strong> and the volume element ∆τ → 0.<br />

Then the Lagrangian can be written as the <strong>in</strong>tegral over the volume element rather than a summation<br />

439


440 CHAPTER 16. ANALYTICAL FORMULATIONS FOR CONTINUOUS SYSTEMS<br />

over ∆. Thatis,<br />

= 1 Z Ã µ ! 2 ( )<br />

˙ 2 − <br />

(16.3)<br />

2<br />

<br />

The discrete-cha<strong>in</strong> coord<strong>in</strong>ate () is assumed to be a cont<strong>in</strong>uous function ( ) for the uniform cha<strong>in</strong>. Thus<br />

the <strong>in</strong>tegral form of the Lagrangian can be expressed as<br />

= 1 Z Ã µ ! 2 Z<br />

( )<br />

˙ 2 − <br />

= L (16.4)<br />

2<br />

<br />

where the function L is called the Lagrangian density def<strong>in</strong>ed by<br />

Ã<br />

L ≡ 1 µ ! 2 ( )<br />

˙ 2 − <br />

2<br />

<br />

(16.5)<br />

The variable <strong>in</strong> the Lagrangian density is not a generalized coord<strong>in</strong>ate; it only serves the role of a cont<strong>in</strong>uous<br />

<strong>in</strong>dex played previously by the <strong>in</strong>dex . Forthediscretecase,eachvalueof def<strong>in</strong>ed a different generalized<br />

coord<strong>in</strong>ate . Now for each value of there is a cont<strong>in</strong>uous function ( ) which is a function of both<br />

position and time.<br />

Lagrange’s equations of motion applied to the cont<strong>in</strong>uous Lagrangian <strong>in</strong> equation 164 gives<br />

2 <br />

2 − 2 <br />

=0 (16.6)<br />

2 This is the familiar wave equation <strong>in</strong> one dimension for a longitud<strong>in</strong>al wave on the cont<strong>in</strong>uous cha<strong>in</strong> with a<br />

phase velocity<br />

s<br />

<br />

=<br />

(16.7)<br />

<br />

The cont<strong>in</strong>uous l<strong>in</strong>ear cha<strong>in</strong> also can exhibit transverse modes which have a Lagrangian density were the<br />

Young’s modulus is replaced by the tension <strong>in</strong> the cha<strong>in</strong>, and is replaced by the l<strong>in</strong>ear mass density <br />

of the cha<strong>in</strong>, lead<strong>in</strong>g to a phase velocity for a transverse wave =<br />

q<br />

<br />

.<br />

16.3 The Lagrangian density formulation for cont<strong>in</strong>uous systems<br />

16.3.1 One spatial dimension<br />

In general the Lagrangian density can be a function of ∇ <br />

<br />

and . Itisof<strong>in</strong>terestthatHamilton’s<br />

pr<strong>in</strong>ciple leads to a set of partial differential equations of motion, based on the Lagrangian density, that are<br />

analogous to the Lagrange equations of motion for discrete systems. When deriv<strong>in</strong>g the Lagrangian equations<br />

of motion <strong>in</strong> terms of the Lagrangian density us<strong>in</strong>g Hamilton’s pr<strong>in</strong>ciple, the notation is simplified if the<br />

system is limited to one spatial coord<strong>in</strong>ate In addition, it is convenient to use the compact notation<br />

where the spatial derivative is written 0 ≡ <br />

<br />

<br />

and the time derivative is ˙ ≡<br />

<br />

, and the one-dimensional<br />

Lagrangian density is assumed to be a function L( 0 ˙ ) The appearance of the derivative 0 ≡ <br />

as<br />

an argument of the Lagrange density is a consequence of the cont<strong>in</strong>uous dependence of on . In pr<strong>in</strong>ciple,<br />

higher-order derivatives could occur but they do not arise <strong>in</strong> most problems of physical <strong>in</strong>terest.<br />

Assum<strong>in</strong>g that the one spatial dimension is , then Hamilton’s pr<strong>in</strong>ciple of least action can be expressed<br />

<strong>in</strong> terms of the Lagrangian density as<br />

Z 2<br />

Z 2<br />

Z 2<br />

= ( ˙ ) = L( 0 ˙ ) (16.8)<br />

1<br />

1 1<br />

Follow<strong>in</strong>g the same approach used <strong>in</strong> chapter 52, it is assumed that the stationary path for the action<br />

<strong>in</strong>tegral is described by the function ( ). Def<strong>in</strong>e a neighbor<strong>in</strong>g function us<strong>in</strong>g a parametric representation<br />

( ; ) such that when =0, the extremum function = ( ) yields the stationary action <strong>in</strong>tegral .


16.3. THE LAGRANGIAN DENSITY FORMULATION FOR CONTINUOUS SYSTEMS 441<br />

Assume that an <strong>in</strong>f<strong>in</strong>itessimal fraction of a neighbor<strong>in</strong>g function ( ) is added to the extremum path<br />

( ). Thatis,assume<br />

L(( ; ) 0 ( ; ) ˙( ; )) (16.12)<br />

( ; ) = ( )+( ) (16.9)<br />

0 ( ; ) ≡<br />

( ; ) ( ) ( )<br />

= + = 0 ( )+ 0 ( )<br />

<br />

(16.10)<br />

˙( ; ) ≡<br />

( ; ) ( ) ( )<br />

= + = ˙( )+ ˙( )<br />

<br />

(16.11)<br />

where it is assumed that both the extremum function ( ) and the auxiliary function ( ) are well<br />

behaved functions of and with cont<strong>in</strong>uous first derivatives, and that ( ) =0at ( 1 1 ) and ( 2 2 )<br />

because, for all possible paths, the function ( ; ) must be identical with ( ) at the end po<strong>in</strong>ts of the<br />

path, i.e. ( 1 1 )=( 2 2 )=0.<br />

A parametric family of curves () as a function of the admixture coefficient , is described by the<br />

function<br />

Z 2<br />

Z 2<br />

() =<br />

1 1<br />

Then Hamilton’s pr<strong>in</strong>ciple requires that the action <strong>in</strong>tegral be a stationary function value for =0,thatis,<br />

() is <strong>in</strong>dependent of which is satisfied if<br />

()<br />

<br />

=<br />

Z 2<br />

Z 2<br />

1<br />

1<br />

µ L <br />

<br />

+ L<br />

˙<br />

Equations 169 1610and 1611 give the partial differentials<br />

˙<br />

+ L 0 <br />

0 =0 (16.13)<br />

<br />

<br />

<br />

= ( ) (16.14)<br />

0<br />

<br />

= 0 ( ) (16.15)<br />

˙<br />

<br />

= ˙( ) (16.16)<br />

Integration by parts <strong>in</strong> both the and terms <strong>in</strong> equation 1613 plus us<strong>in</strong>g the fact that ( 1 1 )=<br />

( 2 2 )=0at both end po<strong>in</strong>ts, yields<br />

Z 2<br />

<br />

Z<br />

1<br />

2<br />

Therefore Hamilton’s pr<strong>in</strong>ciple, equation 1613 becomes<br />

()<br />

<br />

=<br />

1<br />

Z 2<br />

Z 2<br />

1<br />

Z<br />

L ˙<br />

2<br />

µ <br />

˙ = − L <br />

(16.17)<br />

1<br />

˙ <br />

L 0 Z 2<br />

µ <br />

0 = − L <br />

0 (16.18)<br />

<br />

1<br />

∙ L<br />

− <br />

1<br />

µ L<br />

− µ ¸ L<br />

˙ 0 ( ) =0 (16.19)<br />

S<strong>in</strong>ce the auxiliary function ( ) is arbitrary, then the <strong>in</strong>tegrand term <strong>in</strong> the square brackets of equation<br />

1619 must equal zero. That is, µ <br />

L<br />

+ µ L<br />

˙ 0 − L =0 (16.20)<br />

<br />

Equation 1620 gives the equations of motion <strong>in</strong> terms of the Lagrangian density that has been derived<br />

based on Hamilton’s pr<strong>in</strong>ciple.<br />

16.3.2 Three spatial dimensions<br />

Equation 164 expresses the Lagrangian as an <strong>in</strong>tegral of the Lagrangian density over a s<strong>in</strong>gle cont<strong>in</strong>uous<br />

<strong>in</strong>dex ( ) where the Lagrangian density is a function L( ). The derivation of the Lagrangian


442 CHAPTER 16. ANALYTICAL FORMULATIONS FOR CONTINUOUS SYSTEMS<br />

equations of motion <strong>in</strong> terms of the Lagrangian density for three spatial dimensions <strong>in</strong>volves the straightforward<br />

addition of the and coord<strong>in</strong>ates. That is, <strong>in</strong> three dimensions the vector displacement is expressed<br />

by the vector q ( ) and the Lagrangian density is related to the Lagrangian by <strong>in</strong>tegration over three<br />

dimensions. That is, they are related by the equation<br />

L<br />

q<br />

<br />

Z<br />

=<br />

L<br />

q<br />

<br />

L(q q ∇ · q) (16.21)<br />

<br />

where, <strong>in</strong> cartesian coord<strong>in</strong>ates, the volume element = . The Lagrangian density is a function<br />

L(q q<br />

<br />

∇ · q) where the one field quantity ( ) has been extended to a spatial vector q ( )<br />

and the spatial derivatives 0 have been transformed <strong>in</strong>to ∇ · q. Apply<strong>in</strong>g the method used for the onedimensional<br />

spatial system, to the three-dimensional system, leads to the follow<strong>in</strong>g set of equations of motion<br />

à ! à ! à ! à !<br />

<br />

+ + + − L =0 (16.22)<br />

q<br />

where the spatial derivatives have been written explicitly for clarity.<br />

Note that the equations of motion, equation 1622, treat the spatial and time coord<strong>in</strong>ates symmetrically.<br />

This symmetry between space and time is unchanged by multiply<strong>in</strong>g the spatial and time coord<strong>in</strong>ate by<br />

arbitrary numerical factors. This suggests the possibility of <strong>in</strong>troduc<strong>in</strong>g a four-dimensional coord<strong>in</strong>ate system<br />

L<br />

q<br />

<br />

≡ { }<br />

where the parameter is freely chosen. Us<strong>in</strong>g this 4-dimensional formalism allows equation 1622 to be<br />

written more compactly as<br />

⎛<br />

4X <br />

⎝ L ⎠ − L =0 (16.23)<br />

<br />

q q<br />

<br />

<br />

⎞<br />

As discussed <strong>in</strong> chapter 17 relativistic mechanics treats time and space symmetrically, that is, a fourdimensional<br />

vector q ( ) can be used that treats time and the three spatial dimensions symmetrically<br />

and equally. This four-dimensional space-time formulation allows the first four terms <strong>in</strong> equation 1622 to be<br />

condensed <strong>in</strong>to a s<strong>in</strong>gle term which illustrates the symmetry underly<strong>in</strong>g equation 1623. If the Lagrangian<br />

density is Lorentz <strong>in</strong>variant, and if = then equation 1623 is covariant. Thus the Lagrangian density<br />

formulation is ideally suited to the development of relativistically covariant descriptions of fields.<br />

16.4 The Hamiltonian density formulation for cont<strong>in</strong>uous systems<br />

L<br />

q<br />

<br />

Chapter 163 illustrates, <strong>in</strong> general terms, how field theory can be expressed <strong>in</strong> a Lagrangian formulation<br />

via use of the Lagrange density. It is equally possible to obta<strong>in</strong> a Hamiltonian formulation for cont<strong>in</strong>uous<br />

systems analogous to that obta<strong>in</strong>ed for discrete systems. As summarized <strong>in</strong> chapter 8, the Hamiltonian<br />

and Hamilton’s canonical equations of motion are related directly to the Lagrangian by use of a Legendre<br />

transformation. The Hamiltonian is def<strong>in</strong>ed as be<strong>in</strong>g<br />

≡ X µ <br />

<br />

˙ − (16.24)<br />

˙<br />

<br />

The generalized momentum is def<strong>in</strong>ed to be<br />

≡ <br />

(16.25)<br />

˙ <br />

Equation (1625) allows the Hamiltonian (1624) to be written <strong>in</strong> terms of the conjugate momenta as<br />

( )= X <br />

˙ − ( ˙ )= X <br />

( ˙ − ( ˙ )) (16.26)<br />

where the Lagrangian has been partitioned <strong>in</strong>to the terms for each of the <strong>in</strong>dividual coord<strong>in</strong>ates, that is,<br />

( ˙ )= P ( ˙ ).


16.5. LINEAR ELASTIC SOLIDS 443<br />

In the limit that the coord<strong>in</strong>ates are cont<strong>in</strong>uous, then the summation <strong>in</strong> equation 1626 can be<br />

transformed <strong>in</strong>to a volume <strong>in</strong>tegral over the Lagrangian density L. In addition, a momentum density can be<br />

represented by the vector field π where<br />

π ≡ L<br />

(16.27)<br />

˙q<br />

Then the obvious def<strong>in</strong>ition of the Hamiltonian density H is<br />

Z Z<br />

= H = (π · ˙q−L) (16.28)<br />

where the Hamiltonian density is def<strong>in</strong>ed to be<br />

H =π · ˙q−L (16.29)<br />

Unfortunately the Hamiltonian density formulation does not treat space and time symmetrically mak<strong>in</strong>g<br />

it more difficult to develop relativistically covariant descriptions of fields. Hamilton’s pr<strong>in</strong>ciple can be used<br />

to derive the Hamilton equations of motion <strong>in</strong> terms of the Hamiltonian density analogous to the approach<br />

used to derive the Lagrangian density equations of motion. As described <strong>in</strong> <strong>Classical</strong> <strong>Mechanics</strong> 2 edition<br />

by Goldste<strong>in</strong>, the resultant Hamilton equations of motion for one dimension are<br />

H<br />

<br />

= ˙ (16.30)<br />

H<br />

− H<br />

0 = − ˙ (16.31)<br />

H<br />

= − L<br />

<br />

(16.32)<br />

Note that equation 1631 differs from that for discont<strong>in</strong>uous systems.<br />

16.5 L<strong>in</strong>ear elastic solids<br />

Elasticity is a property of matter where the atomic forces <strong>in</strong> matter act to restore the shape of a solid when<br />

distorted due to the application of external forces. A perfectly elastic material returns to its orig<strong>in</strong>al shape<br />

if the external force produc<strong>in</strong>g the deformation is removed. Materials are elastic when the external forces<br />

do not exceed the elastic limit. Above the elastic limit, solids can exhibit plastic flow and concomitant heat<br />

dissipation. Such non-elastic behavior <strong>in</strong> solids occurs when they are subject to strong external forces.<br />

The discussion of l<strong>in</strong>ear systems, <strong>in</strong> chapters 3 and 14, focussed on one dimensional systems, such as the<br />

l<strong>in</strong>ear cha<strong>in</strong>, where the transverse rigidity of the cha<strong>in</strong> was ignored. An extension of the one-dimensional<br />

l<strong>in</strong>ear cha<strong>in</strong> to two-dimensional membranes, such as a drum sk<strong>in</strong>, is straightforward if the membrane is th<strong>in</strong><br />

enough so that the rigidity of the membrane can be ignored. Elasticity for three-dimensional solids requires<br />

account<strong>in</strong>g for the strong elastic forces exerted aga<strong>in</strong>st any change <strong>in</strong> shape <strong>in</strong> addition to elastic forces<br />

oppos<strong>in</strong>g change <strong>in</strong> volume. The stiffness of solids to changes <strong>in</strong> shape, or volume, is best represented us<strong>in</strong>g<br />

the concepts of stress and stra<strong>in</strong>.<br />

Forces <strong>in</strong> matter can be divided <strong>in</strong>to two classes; (1) body forces, such as gravity, which act on each<br />

volume element, and (2) surface forces which are the forces that act on both sides of any <strong>in</strong>f<strong>in</strong>itessimal<br />

surface element <strong>in</strong>side the solid. Surface forces can have components along the normal to the <strong>in</strong>f<strong>in</strong>itessimal<br />

surface, as well as shear components <strong>in</strong> the plane of the surface element. Typically solids are elastic to both<br />

normal and shear components of the surface forces whereas shear forces <strong>in</strong> liquids and gases lead to fluid<br />

flow plus viscous forces due to energy dissipation. As described below, the forces act<strong>in</strong>g on an <strong>in</strong>f<strong>in</strong>itessimal<br />

surface element are best expressed <strong>in</strong> terms of the stress tensor, while the relative distortion of the shape,<br />

or volume, of the body are best expressed <strong>in</strong> terms of the stra<strong>in</strong>tensor. Themoduliofelasticityrelatethe<br />

ratio of the correspond<strong>in</strong>g stress and stra<strong>in</strong> tensors. The moduli of elasticity are constant <strong>in</strong> l<strong>in</strong>ear elastic<br />

solids and thus the stress is proportional to the stra<strong>in</strong> provid<strong>in</strong>g that the stra<strong>in</strong>s do not exceed the elastic<br />

limit.


444 CHAPTER 16. ANALYTICAL FORMULATIONS FOR CONTINUOUS SYSTEMS<br />

16.5.1 Stress tensor<br />

Consider an <strong>in</strong>f<strong>in</strong>itessimal surface area A of an arbitrary closed volume element <strong>in</strong>side the medium.<br />

The surface area element is def<strong>in</strong>ed as a vector A = ˆn where ˆn is the outward normal to the closed<br />

surface that encloses the volume element. Assume that F is the force element exerted by the outside on<br />

the material <strong>in</strong>side the volume element. The stress tensor T is def<strong>in</strong>ed as the ratio of F and A where the<br />

force vector F is given by the <strong>in</strong>ner product of the stress tensor T and the surface element vector A. That<br />

is,<br />

F = T·A (16.33)<br />

S<strong>in</strong>ce both F and A are vectors, then equation 1633 implies that the stress tensor must be a second-rank<br />

tensor as described <strong>in</strong> appendix , that is, the stress tensor is analogous to the rotation matrix or the <strong>in</strong>ertia<br />

tensor. Note that if F and ˆnA are col<strong>in</strong>ear, then the stress tensor T reduces to the conventional pressure<br />

The general stress tensor equals the momentum flux density and has the dimensions of pressure.<br />

16.5.2 Stra<strong>in</strong> tensor<br />

Forces applied to a solid body can lead to translational, or rotational acceleration, <strong>in</strong> addition to chang<strong>in</strong>g<br />

the shape or volume of the body. Elastic forces do not act when an overall displacement ξ of an <strong>in</strong>f<strong>in</strong>itessimal<br />

volume occurs, such as is <strong>in</strong>volved <strong>in</strong> translational or rotational motion. Elastic forces act to oppose positiondependent<br />

differences <strong>in</strong> the displacement vector ξ, that is, the stra<strong>in</strong> depends on the tensor product ∇ ⊗ ξ.<br />

For an elastic medium, the stra<strong>in</strong> depends only on the applied stress and not on the prior load<strong>in</strong>g history.<br />

Consider that the matter at the location r is subject to an elastic displacement ξ, andsimilarlyata<br />

displaced location r 0 = r+ P <br />

<br />

where are cartesian coord<strong>in</strong>ates. The net relative displacement<br />

between r and r 0 is given by<br />

∙ µ <br />

2 + <br />

<br />

+ ¸<br />

<br />

(16.34)<br />

<br />

2 = X ( + ) 2 − X ( ) 2 = X<br />

<br />

<br />

<br />

Ignor<strong>in</strong>g the second order term <br />

<br />

equation gives that the component of is<br />

= X 1<br />

2<br />

<br />

Def<strong>in</strong>e the elements of the stra<strong>in</strong> tensor to be given by<br />

µ <br />

+ <br />

<br />

(16.35)<br />

<br />

µ <br />

+ <br />

<br />

<br />

= 1 2<br />

(16.36)<br />

then<br />

= X <br />

<br />

(16.37)<br />

Thus the stra<strong>in</strong> tensor σ is a rank-2 tensor def<strong>in</strong>ed as the ratio of the stra<strong>in</strong> vector ξ and the <strong>in</strong>f<strong>in</strong>itessimal<br />

area vector A<br />

ξ = σ·A (16.38)<br />

where the component form of the rank -2 stra<strong>in</strong> tensor is<br />

1 1 1<br />

σ = 1 1 2 3<br />

2 2 2<br />

2<br />

1 2 3<br />

(16.39)<br />

¯<br />

¯¯¯¯¯¯¯<br />

3 3 3<br />

1 2 3<br />

The potential-energy density for l<strong>in</strong>ear elastic forces is quadratic <strong>in</strong> the stra<strong>in</strong> components. That is, it is<br />

of the form<br />

= X 1<br />

2 (16.40)<br />

<br />

where is a rank-4 tensor. No preferential directions rema<strong>in</strong> for a homogeneous isotropic elastic body<br />

which allows for two contractions, thereby reduc<strong>in</strong>g the potential energy density to the <strong>in</strong>ner product<br />

= X 1<br />

2 ( ) 2 (16.41)


16.5. LINEAR ELASTIC SOLIDS 445<br />

16.5.3 Moduli of elasticity<br />

The modulus of elasticity of a body is def<strong>in</strong>ed to be the slope of the stress-stra<strong>in</strong> curve and thus, <strong>in</strong><br />

pr<strong>in</strong>ciple, it is a complicated rank-4 tensor that characterizes the elastic properties of a material. Thus the<br />

general theory of elasticity is complicated because the elastic properties depend on the orientation of the<br />

microscopic composition of the elastic matter. The theory simplifies considerably for homogeneous, isotropic<br />

l<strong>in</strong>ear materials below the elastic limit, where the stra<strong>in</strong> is proportional to the applied stress. That is, the<br />

modulus of elasticity then reduces by contractions to a constant scalar value that depends on the properties<br />

of the matter <strong>in</strong>volved.<br />

The potential energy density for homogeneous, isotropic, l<strong>in</strong>ear material, equation 1641 can be separated<br />

<strong>in</strong>to diagonal and off-diagonal components of the stra<strong>in</strong> tensor. That is,<br />

"<br />

= 1 X ( ) 2 +2 X #<br />

( ) 2 (16.42)<br />

2<br />

<br />

<br />

The diagonal first term is the dilation term which corresponds to changes <strong>in</strong> the volume with no changes<br />

<strong>in</strong> shape. The off-diagonal second term <strong>in</strong>volves the shear terms that correspond to changes of the shape of<br />

the body that also changes the volume. The constants and are Lamé’s moduli of elasticity which are<br />

positive. The various moduli of elasticity, correspond<strong>in</strong>g to different distortions <strong>in</strong> the shape and volume of<br />

any solid body, can be derived from Lamé’s moduli for the material.<br />

The components of the elastic forces can be derived from the gradient of the elastic potential energy,<br />

equation 1642 by use of Gauss’ law plus vector differential calculus. The components of the elastic force,<br />

derived from the stra<strong>in</strong> tensor σ, can be associated with the correspond<strong>in</strong>g components of the stress tensor<br />

T. Thus, for homogeneous isotropic l<strong>in</strong>ear materials, the components of the stress tensor are related to the<br />

stra<strong>in</strong> tensor by the relation<br />

X<br />

µ<br />

<br />

= <br />

+ + <br />

X<br />

<br />

= +2 (16.43)<br />

<br />

<br />

where it has been assumed that = . The two moduli of elasticity and are material-dependent<br />

constants. Equation 1643 canbewritten<strong>in</strong>tensornotationas<br />

<br />

T = (σ)I +2σ (16.44)<br />

where () isthetraceofthestra<strong>in</strong>tensorand is the identity matrix.<br />

Equation 1644 can be <strong>in</strong>verted to give the stra<strong>in</strong> tensor components <strong>in</strong> terms of the stress tensor components.<br />

"<br />

= 1 −<br />

2<br />

<br />

(3 +2)<br />

#<br />

X<br />

<br />

<br />

(16.45)<br />

The various moduli of elasticity relate comb<strong>in</strong>ations of different stress and stra<strong>in</strong> tensor components. The<br />

follow<strong>in</strong>g five elastic moduli are used frequently to describe elasticity <strong>in</strong> homogeneous isotropic media, and<br />

all are related to Lamé’s two moduli of elasticity.<br />

1) Young’s modulus describes tensile elasticity which is axial stiffness of the length of a body to<br />

deformation along the axis of the applied tensile force.<br />

≡ 11 (3 +2)<br />

=<br />

11 ( + )<br />

(16.46)<br />

2) Bulk modulus = ∆<br />

<br />

def<strong>in</strong>es the relative dilation or compression of a bodies volume to pressure<br />

applied uniformly <strong>in</strong> all directions.<br />

= + 2 3 (16.47)<br />

The bulk modulus is an extension of Young’s modulus to three dimensions and typically is larger than .<br />

The <strong>in</strong>verse of the bulk modulus is called the compressibility of the material.


446 CHAPTER 16. ANALYTICAL FORMULATIONS FOR CONTINUOUS SYSTEMS<br />

3) Shear modulus describes the shear stiffness of a body to volume-preserv<strong>in</strong>g shear deformations.<br />

The shear stra<strong>in</strong> becomes a deformation angle given by the ratio of the displacement along the axis of the<br />

shear force and the perpendicular moment arm. The shear modulus equals Lamé’s constant . Thatis,<br />

= (16.48)<br />

4) Poisson’s ratio is the negative ratio of the transverse to axial stra<strong>in</strong>. It is a measure of the volume<br />

conserv<strong>in</strong>g tendency of a body to contract <strong>in</strong> the directions perpendicular to the axis along which it is<br />

stretched. In terms of Lamé’s constants, Poisson’s ratio equals<br />

<br />

=<br />

(16.49)<br />

2( + )<br />

Note that for a stable, isotropic elastic material, Poisson’s ratio is bounded between −10 ≤ ≤ 05 to ensure<br />

that the and moduli have positive values. At the <strong>in</strong>compressible limit, =05, and the bulk modulus<br />

and Lame parameter are <strong>in</strong>f<strong>in</strong>ite, that is, the compressibility is zero. Typical solids have Poisson’s ratios<br />

of ≈ 005 if hard and =025 if soft.<br />

The stiffness of elastic solids <strong>in</strong> terms of the elastic moduli of solids can be complicated due to the<br />

geometry and composition of solid bodies. Often it is more convenient to express the stiffness <strong>in</strong> terms of<br />

the spr<strong>in</strong>g constant where<br />

= <br />

(16.50)<br />

<br />

The spr<strong>in</strong>g constant is <strong>in</strong>versely proportional to the length of the spr<strong>in</strong>g because the stra<strong>in</strong> of the material<br />

is def<strong>in</strong>ed to be the fractional deformation, not the absolute deformation.<br />

16.5.4 Equations of motion <strong>in</strong> a uniform elastic media<br />

The divergence theorem (8) relates the volume <strong>in</strong>tegral of the divergence of T to the vector force density<br />

F act<strong>in</strong>g on the closed surface. I Z<br />

Z<br />

F = T·A = ∇ · T = f (16.51)<br />

That is, the <strong>in</strong>ner product of the del operator, ∇, I andtherank-2 stress tensor T, givethevectorforce<br />

density f. This force act<strong>in</strong>g on the enclosed mass for the closed volume, leads to an acceleration 2 <br />

<br />

. 2<br />

Thus<br />

I Z<br />

I<br />

F = T·A = ∇ · T = 2 ξ<br />

(16.52)<br />

2 Use equation 1644 to relate the stress tensor T to the moduli of elasticity gives<br />

2 ξ <br />

2 = X "<br />

#<br />

2 ξ <br />

( + ) + 2 ξ <br />

<br />

<br />

2 (16.53)<br />

<br />

where =1 2 3. In general this equation is difficult to solve. However, for the simple case of a plane wave<br />

<strong>in</strong> the =1direction, the problem reduces to the follow<strong>in</strong>g three equations<br />

2 ξ 1<br />

2 = ( +2) 2 ξ 1<br />

2 1<br />

2 ξ 2<br />

2 = 2 ξ 2<br />

2 1<br />

2 ξ 3<br />

2 = 2 ξ 3<br />

2 1<br />

(16.54)<br />

(16.55)<br />

(16.56)<br />

q<br />

Equation 1654 corresponds to a longitud<strong>in</strong>al wave travell<strong>in</strong>g with velocity = (+2)<br />

<br />

. Equations<br />

q<br />

1655 1656 correspond to two perpendicular transverse waves travell<strong>in</strong>g with velocity = <br />

<br />

. This illustrates<br />

the important fact that longitud<strong>in</strong>al waves travel faster than transverse waves <strong>in</strong> an elastic solid.<br />

Seismic waves <strong>in</strong> the Earth, generated by earthquakes, exhibit this property. Note that shear<strong>in</strong>g stresses do<br />

not exist <strong>in</strong> ideal liquids and gases s<strong>in</strong>ce they cannot ma<strong>in</strong>ta<strong>in</strong> shear forces and thus =0


16.6. ELECTROMAGNETIC FIELD THEORY 447<br />

16.6 Electromagnetic field theory<br />

16.6.1 Maxwell stress tensor<br />

Analytical formulations for cont<strong>in</strong>uous systems, developed for describ<strong>in</strong>g elasticity, are generally applicable<br />

when applied to other fields, such as the electromagnetic field. The use of the Maxwell’s stress tensor T to<br />

describe momentum <strong>in</strong> the electromagnetic field, is an important example of the application of cont<strong>in</strong>uum<br />

mechanics <strong>in</strong> field theory.<br />

TheLorentzforcecanbewrittenas<br />

Z<br />

Z<br />

Z<br />

F = (E + v × B) = (E + J × B) = f (16.57)<br />

where the force density f is def<strong>in</strong>ed to be<br />

Maxwell’s equations<br />

= 0 ∇ · E<br />

f =(E + J × B) (16.58)<br />

J = 1 0<br />

∇ × B − ² 0<br />

E<br />

<br />

(16.59)<br />

can be used to elim<strong>in</strong>ate the charge and current densities <strong>in</strong> equation 1657<br />

µ <br />

1 E<br />

f = 0 (∇ · E) E + ∇ × B − ² 0 ×B (16.60)<br />

0 <br />

Vector calculus gives that<br />

<br />

(E × B) =E<br />

× B + E×B<br />

(16.61)<br />

<br />

while Faraday’s law gives<br />

B<br />

= −∇ × E (16.62)<br />

<br />

Equation 1662 allows equation 1661 to be rewritten as<br />

E<br />

× B =+ (E × B) − E×B<br />

=+ (E × B)+E× (∇ × E)<br />

<br />

(16.63)<br />

Equation 1663 can be <strong>in</strong>serted <strong>in</strong>to equation 1660. In addition, a term 1 <br />

(∇ · B) B canbeaddeds<strong>in</strong>ce<br />

0<br />

∇ · B =0 which allows equation 1660 tobewritten<strong>in</strong>thesymmetricform<br />

f = 0 (∇ · E) E + 1 (∇ · B) B+ 1 E<br />

(∇ × B) × B − ² 0 ×B<br />

0 0 <br />

(16.64)<br />

= 0 (∇ · E) E + 1 (∇ · B) B+ 1 <br />

(∇ × B) × B− 0<br />

0 0 (E × B) − 0E× (∇ × E) (16.65)<br />

Us<strong>in</strong>g the vector identity<br />

∇ (A · B) =A× (∇ × B)+B× (∇ × A)+(A · ∇) B+(B · ∇) A (16.66)<br />

Let A = B = E then<br />

That is<br />

Similarly<br />

∇ ¡ 2¢ =2E× (∇ × E)+2(E · ∇) E (16.67)<br />

E× (∇ × E) = 1 2 ∇ ¡ 2¢ − (E · ∇) E (16.68)<br />

B× (∇ × B) = 1 2 ∇ ¡ 2¢ − (B · ∇) B (16.69)<br />

Insert<strong>in</strong>g equations 1668 and 1669 <strong>in</strong>to equation 1665 gives<br />

∙<br />

f= 0 (∇ · E) E+(E · ∇) E− 1 ∇2¸<br />

+ 1 ∙<br />

(∇ · B) B+(B · ∇) B− 1 ∇2¸ <br />

− 0 (E × B) (16.70)<br />

2 0 2


448 CHAPTER 16. ANALYTICAL FORMULATIONS FOR CONTINUOUS SYSTEMS<br />

This complicated formula can be simplified by def<strong>in</strong><strong>in</strong>g the rank-2 Maxwell stress tensor T which has<br />

components<br />

µ<br />

≡ 0 − 1 <br />

2 2 + 1 µ<br />

− 1 <br />

0 2 2 (16.71)<br />

The <strong>in</strong>ner product of the del operator and the Maxwell stress tensor is a vector with components of<br />

∙<br />

(∇ · T) <br />

= 0 (∇ · E) +(E · ∇) − 1 2¸<br />

2 ∇2 + 1 ∙<br />

(∇ · B) +(B · ∇) − 1 2¸<br />

0 2 ∇2 (16.72)<br />

The above def<strong>in</strong>ition of the Maxwell stress tensor, plus the Poynt<strong>in</strong>g vector S = 1 <br />

(E × B) allows the force<br />

0<br />

density equation 1658 tobewritten<strong>in</strong>theform<br />

S<br />

f = ∇ · T− 0 0 (16.73)<br />

<br />

The divergence theorem allows the total force, act<strong>in</strong>g of the volume to be written <strong>in</strong> the form<br />

Z µ<br />

<br />

S<br />

F = ∇ · T− 0 0 (16.74)<br />

<br />

I<br />

Z<br />

<br />

= T·a− 0 0 Sdτ (16.75)<br />

<br />

Note that, if the Poynt<strong>in</strong>g vector is time <strong>in</strong>dependent, then the second term <strong>in</strong> equation 1675 is zero and the<br />

Maxwell stress tensor T is the force per unit area, (stress) act<strong>in</strong>g on the surface. The fact that T is a rank-2<br />

tensor is apparent s<strong>in</strong>ce the stress represents the ratio of the force-density vector f and the <strong>in</strong>f<strong>in</strong>itessimal<br />

area vector a, which do not necessarily po<strong>in</strong>t <strong>in</strong> the same directions.<br />

16.6.2 Momentum <strong>in</strong> the electromagnetic field<br />

Chapter 72 showed that the electromagnetic field carries a l<strong>in</strong>ear momentum A where isthechargeona<br />

body and A is the electromagnetic vector potential. It is useful to use the Maxwell stress tensor to express<br />

the momentum density directly <strong>in</strong> terms of the electric and magnetic fields.<br />

Newton’s law of motion can be used to write equation equation 1675 as<br />

F= p <br />

<br />

I<br />

=<br />

T·a− 0 0<br />

<br />

<br />

Z<br />

Sdτ (16.76)<br />

where p is the total mechanical l<strong>in</strong>ear momentum of the volume . Equation1676 implies that the electromagnetic<br />

field carries a l<strong>in</strong>ear momentum<br />

p = 0 0<br />

Z<br />

Sdτ (16.77)<br />

I<br />

The T·a term <strong>in</strong> equation 1676 is the momentum per unit time flow<strong>in</strong>g <strong>in</strong>to the closed surface.<br />

In field theory it can be useful to describe the behavior <strong>in</strong> terms of the momentum flux density π. Thus<br />

the momentum flux density π <strong>in</strong> the electromagnetic field is<br />

π = 0 0 S (16.78)<br />

Then equation 1676 implies that the total momentum flux density π = π +π is related to Maxwell’s<br />

stress tensor by<br />

<br />

(π + π )=∇ · T (16.79)<br />

That is, like the elasticity stress tensor, the divergence of Maxwell’s stress tensor T equalstherateofchange<br />

of the total momentum density, that is, −T is the momentum flux density.<br />

This discussion of the Maxwell stress tensor and its relation to momentum <strong>in</strong> the electromagnetic field<br />

illustrates the role that analytical formulations of classical mechanics can play <strong>in</strong> field theory.


16.7. IDEAL FLUID DYNAMICS 449<br />

16.7 Ideal fluid dynamics<br />

The dist<strong>in</strong>ction between a solid and a fluid is that a fluid flows under shear stress whereas the elasticity<br />

of solids oppose distortion and flow. Shear stress <strong>in</strong> a fluid is opposed by dissipative viscous forces, which<br />

depend on velocity, as opposed to elastic solids where the shear stress is opposed by the elastic forces which<br />

depend on the displacement. An ideal fluid is one where the viscous forces are negligible, and thus the shear<br />

stress Lamé parameter =0.<br />

16.7.1 Cont<strong>in</strong>uity equation<br />

Fluid dynamics requires a different philosophical approach than that used to describe the motion of an<br />

ensemble of known solid bodies.The prior discussions of classical mechanics used, as variables, the coord<strong>in</strong>ates<br />

of each member of an ensemble of particles with known masses. This approach is not viable for fluids<br />

which <strong>in</strong>volve an enormous number of <strong>in</strong>dividual atoms as the fundamental bodies of the fluid. The best<br />

philosophical approach for describ<strong>in</strong>g fluid dynamics is to employ cont<strong>in</strong>uum mechanics us<strong>in</strong>g def<strong>in</strong>ite fixed<br />

volume elements and describe the fluid <strong>in</strong> terms of macroscopic variables of the fluid such as mass density<br />

, pressure ,andfluid velocity v.<br />

Conservation of fluid mass requires that the rate of change of mass <strong>in</strong> a fixed volume must equal the net<br />

<strong>in</strong>flow of mass.<br />

Z I<br />

<br />

+ v·a = 0 (16.80)<br />

<br />

Us<strong>in</strong>g the divergence theorem (2) allows this to be written as<br />

Z µ <br />

<br />

+ ∇· (v) =0 (16.81)<br />

<br />

Mass conservation must hold for any arbitrary volume, therefore the cont<strong>in</strong>uity equation can be written <strong>in</strong><br />

the differential form<br />

<br />

+ ∇· (v) =0 (16.82)<br />

<br />

16.7.2 Euler’s hydrodynamic equation<br />

The fluid surround<strong>in</strong>g a volume exerts a net force F that equals the surface <strong>in</strong>tegral of the pressure P.<br />

This force can be transformed to a volume <strong>in</strong>tegral of ∇ . The net force then will lead to an acceleration<br />

of the volume element. That is<br />

I<br />

F = −<br />

Z<br />

a = −<br />

Z<br />

∇ =<br />

v (16.83)<br />

<br />

Thus the force density f is given by<br />

f = −∇P = v<br />

(16.84)<br />

<br />

Note that the acceleration v<br />

<br />

<strong>in</strong> equation 1683 referstotherateofchangeofvelocityfor<strong>in</strong>dividual<br />

atoms <strong>in</strong> the fluid, nottherateofchangeoffluid velocity at a fixed po<strong>in</strong>t <strong>in</strong> space. These two accelerations<br />

are related by not<strong>in</strong>g that, dur<strong>in</strong>g the time , the change <strong>in</strong> velocity v of a given fluid particle is composed<br />

of two parts, namely (1) the change dur<strong>in</strong>g <strong>in</strong> the velocity at a fixed po<strong>in</strong>t <strong>in</strong> space, and (2) the difference<br />

between the velocities at that same <strong>in</strong>stant <strong>in</strong> time at two po<strong>in</strong>ts displaced a distance r apart, where r is<br />

thedistancemovedbyagivenfluid particle dur<strong>in</strong>g the time . The firstpartisgivenby v<br />

<br />

at a given<br />

po<strong>in</strong>t ( ) <strong>in</strong> space. The second part equals<br />

Thus<br />

v v v<br />

+ + =(r · ∇) v (16.85)<br />

<br />

v = v +(r · ∇) v (16.86)


450 CHAPTER 16. ANALYTICAL FORMULATIONS FOR CONTINUOUS SYSTEMS<br />

Divide both sides by gives that the acceleration of the atoms <strong>in</strong> the fluid equals<br />

Substitute equation 1687 <strong>in</strong>to 1684 gives<br />

v<br />

= v +(v · ∇) v (16.87)<br />

<br />

v<br />

+(v · ∇) v = − 1 ∇ (16.88)<br />

<br />

This is Euler’s equation for hydrodynamics. The two terms on the left represent the acceleration <strong>in</strong> the<br />

<strong>in</strong>dividual fluid components while the right-hand side lists the force density produc<strong>in</strong>g the acceleration.<br />

Additional forces can be added to the right-hand side. For example, the gravitational force density g<br />

can be expressed <strong>in</strong> terms of the gravitational scalar potential to be<br />

Inclusion of the gravitational field force density <strong>in</strong> Euler’s equation gives<br />

16.7.3 Irrotational flow and Bernoulli’s equation<br />

g = −ρ∇ (16.89)<br />

v<br />

+(v · ∇) v = − 1 ∇ ( + ) (16.90)<br />

<br />

Streaml<strong>in</strong>ed flow corresponds to irrotational flow, that is, ∇ × v = 0. S<strong>in</strong>ce irrotational flow is curl free, the<br />

velocity streaml<strong>in</strong>es can be represented by a scalar potential field . Thatis<br />

v = −∇ (16.91)<br />

This scalar potential field can be used to derive the vector velocity field for irrotational flow.<br />

Note that the (v · ∇) v term <strong>in</strong> Euler’s equation (1690) can be rewritten us<strong>in</strong>g the vector identity<br />

(v · ∇) v = 1 2 ∇ ¡ 2¢ − v × ∇ × v (16.92)<br />

Insert<strong>in</strong>g equation 1692 <strong>in</strong>to Euler’s equation 1690 then gives<br />

v<br />

= v × ∇ × v− 1 ∇ µ 1<br />

2 2 + + <br />

Potential flow corresponds to time <strong>in</strong>dependent irrotational flow, that is, both v<br />

<br />

potential flow equation 1693 reduces to<br />

µ <br />

1<br />

∇<br />

2 2 + + =0<br />

<br />

(16.93)<br />

=0and ∇ × v =0 For<br />

which implies that µ <br />

1<br />

2 2 + + = constant (16.94)<br />

This is the famous Bernoulli’s equation that relates the <strong>in</strong>terplay of the fluid velocity, pressure and gravitational<br />

energy. Bernoulli’s equation plays important roles <strong>in</strong> both hydrodynamics and aerodynamics.<br />

16.7.4 Gas flow<br />

Fluid dynamics applied to gases is a straightforward extension of fluid dynamics that employs standard thermodynamical<br />

concepts. The follow<strong>in</strong>g example illustrates the application of fluid mechanics for calculat<strong>in</strong>g<br />

the velocity of sound <strong>in</strong> a gas.


16.7. IDEAL FLUID DYNAMICS 451<br />

16.1 Example: Acoustic waves <strong>in</strong> a gas<br />

Propagation of acoustic waves <strong>in</strong> a gas provides an example of us<strong>in</strong>g the three-dimensional Lagrangian<br />

density. Only longitud<strong>in</strong>al waves occur <strong>in</strong> a gas and the velocity is given by thermodynamics of the gas. Let<br />

the displacement of each gas molecule be designated by the general coord<strong>in</strong>ate q with correspond<strong>in</strong>g velocity<br />

˙q. Let the gas density be then the k<strong>in</strong>etic energy density () of an <strong>in</strong>f<strong>in</strong>itessimal volume of gas ∆ is<br />

given by<br />

∆ ()= 1 2 0 ˙q 2<br />

The rapid contractions and expansions of the gas <strong>in</strong> an acoustic wave occur adiabatically such that the product<br />

specific heatatconstantpressure<br />

is a constant, where =<br />

specific heatatconstantvolume<br />

. Therefore the change <strong>in</strong> potential energy density<br />

∆() is given to second order by<br />

∆ ()= 1 Z 0 +∆<br />

= 0<br />

∆ + 1 µ <br />

(∆) 2 = 0<br />

∆ − 1 µ<br />

<br />

0<br />

(∆) 2<br />

0 0<br />

0 2 0 <br />

0<br />

0 2 0 0<br />

S<strong>in</strong>ce the volume and density are related by<br />

= 0<br />

then the fractional change <strong>in</strong> the density is related to the density by<br />

= 0 (1 + )<br />

This implies that the potential energy density () is given by<br />

∙<br />

∆ ()= 0 + 2¸<br />

0<br />

2<br />

The mass flow<strong>in</strong>g out of the volume 0 must equal the fractional change <strong>in</strong> density of the volume, that is<br />

Z<br />

Z<br />

0 q · dS = −ρ 0 <br />

The divergence theorem gives that<br />

Z<br />

Z<br />

q · dS =<br />

Z<br />

∇ · q = −<br />

<br />

Thus the density is given by m<strong>in</strong>us the divergence of q<br />

= −∇ · q<br />

This allows the potential energy density to be written as<br />

∆()=− 0 ∇ · q+ 0<br />

2<br />

(∇ · q)2<br />

Comb<strong>in</strong><strong>in</strong>g the k<strong>in</strong>etic energy density and the potential energy density gives the complete Lagrangian density<br />

foranacousticwave<strong>in</strong>agastobe<br />

L = 1 2 0 ˙q 2 + 0 ∇ · q− 0<br />

2<br />

(∇ · q)2<br />

Insert<strong>in</strong>g this Lagrangian density <strong>in</strong> the correspond<strong>in</strong>g equations of motion, equation 1623, givesthat<br />

∇ 2 q− 0 2 q<br />

0 2 =0<br />

where 0 and 0 are the ambient pressure and density of the gas. This is the wave equation where the phase<br />

velocity of sound is given by<br />

s<br />

0<br />

=<br />

0


452 CHAPTER 16. ANALYTICAL FORMULATIONS FOR CONTINUOUS SYSTEMS<br />

16.8 Viscous fluid dynamics<br />

Viscous fluid dynamics is a branch of classical mechanics that plays a pivotal role <strong>in</strong> a wide range of aspects<br />

of life, such as blood flow <strong>in</strong> human anatomy, weather, hydraulic eng<strong>in</strong>eer<strong>in</strong>g, and transportation by land,<br />

sea, and air. Viscous fluid flow provides natures most common manifestation of nonl<strong>in</strong>earity and turbulence<br />

<strong>in</strong> classical mechanics, and provides an excellent illustration of possible solutions of non-l<strong>in</strong>ear equations of<br />

motion <strong>in</strong>troduced <strong>in</strong> chapter 4. A detailed description of turbulence rema<strong>in</strong>s a challeng<strong>in</strong>g problem and<br />

this subject has the reputation of be<strong>in</strong>g the last great unsolved problem <strong>in</strong> classical mechanics. There is<br />

an apocryphal story that Werner Heisenberg was asked, if given the opportunity, what would he like to ask<br />

God. His reply was “When I meet God, I am go<strong>in</strong>g to ask him two questions: Why relativity? and why<br />

turbulence?, I really believe he will only have an answer to the first”.<br />

In contrast to solids, fluids do not have elastic restor<strong>in</strong>g forces to support shear stress because the fluid<br />

flows. Shear stresses <strong>in</strong> fluids are balance by viscous forces which are velocity dependent. There are two<br />

mechanisms that lead to shear stress act<strong>in</strong>g between adjacent fluid layers <strong>in</strong> relative motion. The first<br />

mechanism <strong>in</strong>volves lam<strong>in</strong>ar flow where the viscous forces produce shear stress between adjacent layers of<br />

the fluid which are mov<strong>in</strong>g parallel along adjacent streaml<strong>in</strong>es at differ<strong>in</strong>g velocities. Viscous forces typically<br />

dom<strong>in</strong>ate lam<strong>in</strong>ar flow. High viscosity fluids like honey exhibit lam<strong>in</strong>ar flow and are more difficult to stir<br />

or pour compared with low-viscosity fluids like water. The second mechanism <strong>in</strong>volves turbulent flow where<br />

shear stress is due to momentum transfer between adjacent layers when the flow breaks up <strong>in</strong>to large-scale<br />

coherent vortex structures which carry most of the k<strong>in</strong>etic energy. These eddies lead to transverse motion<br />

that transfers momentum plus heat between adjacent layers and leads to higher drag. The w<strong>in</strong>g-tip vortex<br />

produced by the w<strong>in</strong>g tip of an aircraft is an example of a dynamically-dist<strong>in</strong>ct, large-scale, coherent vortex<br />

structure which has considerable angular momentum and decays by fragmentation <strong>in</strong>to a cascade of smaller<br />

scale structures.<br />

16.8.1 Navier-Stokes equation<br />

Viscous forces act<strong>in</strong>g on the small-scale coherent structures eventually dissipate the energy <strong>in</strong> turbulent<br />

motion. The viscous drag can be handled <strong>in</strong> terms of a stress tensor T analogous to its use when account<strong>in</strong>g<br />

for the elastic restor<strong>in</strong>g forces <strong>in</strong> elasticity as discussed <strong>in</strong> chapter 1653. That is, the viscous force density<br />

is related to the deceleration of the volume element by<br />

<br />

(v) =−∇ · T (16.95)<br />

<br />

where the components of the stress tensor are<br />

= = + (16.96)<br />

Note that the stress tensor gives the momentum flux density tensor, which <strong>in</strong>volves a diagonal term proportional<br />

to pressure plus a viscous drag term that is is proportional to the product of two velocities.<br />

The Navier-Stokes equations are the fundamental equations characteriz<strong>in</strong>g fluid flow. They are based on<br />

application of Newton’s second law of motion to fluids together with the assumption that the fluid stress<br />

is the sum of a diffus<strong>in</strong>g viscous term plus a pressure term. Comb<strong>in</strong><strong>in</strong>g Euler’s equation, 1690, with1695<br />

gives the Navier-Stokes equation<br />

∙ ∇v¸<br />

v<br />

<br />

+ v · = −∇ + ∇ · T+f (16.97)<br />

where is the fluid density, v is the flow velocity vector, the pressure, T is the shear stress tensor viscous<br />

drag term, and f represents external body forces per unit volume such as gravity act<strong>in</strong>g on the fluid.<br />

For <strong>in</strong>compressible flow the stress tensor term simplifies to ∇ · T =∇ 2 v. Then the Navier-Stokes<br />

equation simplifies to<br />

∙ ∇v¸<br />

v<br />

<br />

+ v · = −∇ + ∇ 2 v+f (16.98)<br />

where ∇ 2 v is the viscosity drag term. The left-hand side of equation 1698 represents the rate of change<br />

of momentum per unit volume while the right-hand side represents the summation of the forces per unit<br />

volume that are act<strong>in</strong>g.


16.8. VISCOUS FLUID DYNAMICS 453<br />

The Navier-Stokes equations are nonl<strong>in</strong>ear due to the (v · ∇) v term as well as be<strong>in</strong>g a function of<br />

velocity. This non-l<strong>in</strong>earity leads to a wide spectrum of dynamic behavior rang<strong>in</strong>g from ordered lam<strong>in</strong>ar<br />

flow to chaotic turbulence. Numerical solution of the Navier-Stokes equations is extremely difficult because<br />

of the wide dynamic range of the dimensions of the coherent structures <strong>in</strong>volved <strong>in</strong> turbulent motion. For<br />

example, simulation calculations require use of a high resolution mesh which is a challenge to the capabilities<br />

of current generation computers.<br />

The microscopic boundary condition at the <strong>in</strong>terface of the solid and fluid is that the fluid molecules<br />

have zero average tangential velocity relative to the normal to the solid-fluid <strong>in</strong>terface. This implies that<br />

there is a boundary layer for which there is a gradient <strong>in</strong> the tangential velocity of the fluid between the<br />

solid-fluid <strong>in</strong>terface and the free-steam velocity. This velocity gradient produces vorticity <strong>in</strong> the fluid. When<br />

the viscous forces are negligible then the angular momentum <strong>in</strong> any coherent vortex structure is conserved<br />

lead<strong>in</strong>g to the vortex motion be<strong>in</strong>g preserved as it propagates.<br />

16.8.2 Reynolds number<br />

Fluid flow can be characterized by the Reynolds number<br />

Re which is a dimensionless number that is a measure<br />

of the ratio of the <strong>in</strong>ertial forces 2 to viscous forces<br />

2 .Thatis,<br />

Inertial forces<br />

Re ≡<br />

Viscous forces = <br />

<br />

= (16.99)<br />

<br />

where is the relative velocity between the free fluid<br />

flow and the solid surface, is a characteristic l<strong>in</strong>ear<br />

dimension, is the dynamic viscosity of the fluid, is<br />

the k<strong>in</strong>ematic viscosity ( = <br />

),and is the density<br />

of the fluid. The Law of Similarity implies that at a<br />

given Reynolds number, for a specific shaped solid body,<br />

the fluid flow behaves identically <strong>in</strong>dependent of the size<br />

of the body. Thus one can use small models <strong>in</strong> w<strong>in</strong>d<br />

tunnels, or water-flow tanks, to accurately model fluid<br />

flow that can be scaled up to a full-sized aircraft or boats<br />

by scal<strong>in</strong>g and to give the same Reynolds number.<br />

C D<br />

10 -1 10 0 10 1 10 2 10 3 10 4 10 5<br />

100<br />

A<br />

10<br />

B<br />

1<br />

C<br />

D<br />

0.1<br />

E<br />

10 6 10 7<br />

Reynolds number<br />

No separation<br />

(A)<br />

Steady separation bubble<br />

(B)<br />

16.8.3 Lam<strong>in</strong>ar and turbulent fluid flow<br />

Fluid flow over a cyl<strong>in</strong>der illustrates the general features<br />

of fluid flow. The drag force act<strong>in</strong>g on a cyl<strong>in</strong>der<br />

of diameter and length with the cyl<strong>in</strong>drical axis<br />

perpendicular to the fluid flow, is given by<br />

= 1 2 2 (16.100)<br />

where is the coefficient of drag. Figure 161<br />

shows the dependence of the drag coefficient as a<br />

function of the Reynolds number, for fluid flow that<br />

is transverse to a smooth circular cyl<strong>in</strong>der. The lower<br />

part of figure 161 shows the streaml<strong>in</strong>es for flow around<br />

the cyl<strong>in</strong>der at various Reynolds numbers for the po<strong>in</strong>ts<br />

identified by the letters , and on the plot<br />

of the drag coefficient versus Reynolds number for a<br />

smooth cyl<strong>in</strong>der.<br />

A) At low velocities, where Re ≤ 1 the flow is lam<strong>in</strong>ar<br />

around the cyl<strong>in</strong>der <strong>in</strong> that the low vorticity is<br />

damped by the viscous forces and the v<br />

<br />

term <strong>in</strong> equation<br />

1698 can be ignored. The coefficient of drag <br />

Oscillat<strong>in</strong>g Karman vortex street wake<br />

Lam<strong>in</strong>ar boundary layer,<br />

wide turbulent wake<br />

(D)<br />

(C)<br />

Turbulent boundary layer,<br />

narrow turbulent wake<br />

Figure 16.1: Upper: The dependence of the coefficient<br />

of drag on Reynolds number Re for fluid<br />

flow perpendicular to a smooth circular cyl<strong>in</strong>der<br />

of diameter and length . Lower: Typical flow<br />

patterns for flow past a circular cyl<strong>in</strong>der at various<br />

Reynolds numbers as <strong>in</strong>dicated <strong>in</strong> the upper<br />

figure.<br />

(E)


454 CHAPTER 16. ANALYTICAL FORMULATIONS FOR CONTINUOUS SYSTEMS<br />

varies <strong>in</strong>versely with Re lead<strong>in</strong>g to the drag forces that are roughly l<strong>in</strong>ear with velocity as described <strong>in</strong> chapter<br />

2105 The size and velocities of ra<strong>in</strong>drops <strong>in</strong> a light ra<strong>in</strong> shower correspond to such Reynolds numbers.<br />

B) For 10 Re 30 the flow has two turbulent vortices immediately beh<strong>in</strong>d the body <strong>in</strong> the wake of<br />

the cyl<strong>in</strong>der, but the flow still is primarily lam<strong>in</strong>ar as illustrated.<br />

C) For 40 Re 250 the pair of vortices peel off alternately produc<strong>in</strong>g a regular periodic sequence of<br />

vortices although the flow still is lam<strong>in</strong>ar. This vortex sheet is called a von Kármán vortex sheet for which<br />

the velocity at a given position, relative to the cyl<strong>in</strong>der, is time dependent <strong>in</strong> contrast to the situation at<br />

lower Reynolds numbers.<br />

D) For 10 3 Re 10 5 viscous forces are negligible relative to the <strong>in</strong>ertial effects of the vortices and<br />

boundary-layer vortices have less time to diffuse <strong>in</strong>to the larger region of the fluid, thus the boundary layer is<br />

th<strong>in</strong>ner. The boundary-layer flow exhibits a small scale chaotic turbulence <strong>in</strong> three dimensions superimposed<br />

on regular alternat<strong>in</strong>g vortex structures. In this range is roughly constant and thus the drag forces are<br />

proportional to the square of the velocity. This regime of Reynold numbers corresponds to typical velocities<br />

of mov<strong>in</strong>g automobiles.<br />

E) For Re ≈ 10 6 , which is typical of a fly<strong>in</strong>g aircraft, the <strong>in</strong>ertial effects dom<strong>in</strong>ate except <strong>in</strong> the narrow<br />

boundary layer close to the solid-fluid <strong>in</strong>terface. The chaotic region works its way further forward on the<br />

cyl<strong>in</strong>der reduc<strong>in</strong>g the volume of the chaotic turbulent boundary layer which results <strong>in</strong> a significant decreases<br />

<strong>in</strong> . For a sailplane w<strong>in</strong>g fly<strong>in</strong>g at about 50, the boundary layer at the lead<strong>in</strong>g edge of the cyl<strong>in</strong>der<br />

reduces to the order of a millimeter <strong>in</strong> thickness at the lead<strong>in</strong>g edge and a centimeter at the trail<strong>in</strong>g edge. At<br />

these Reynold’s numbers the airflow comprises a th<strong>in</strong> boundary layer, where viscous effects are important,<br />

plus fluid flow <strong>in</strong> the bulk of the fluid where the vortex <strong>in</strong>ertial terms dom<strong>in</strong>ate and viscous forces can be<br />

ignored. That is, the viscous stress tensor term ∇ · T on the right-hand side of equation 1697 can be<br />

ignored, and the Navier-Stokes equation reduces to the simpler Euler equation for such <strong>in</strong>viscid fluid flow.<br />

The importance of the <strong>in</strong>ertia of the vortices is illustrated by the persistence of the vortex structure<br />

and turbulence over a wide range of length scales characteristic of turbulent flow. The dynamic range of<br />

the dimension of coherent vortex structures is enormous. For example, <strong>in</strong> the atmosphere the vortex size<br />

ranges from 10 5 <strong>in</strong> diameter for hurricanes down to 10 −3 <strong>in</strong> th<strong>in</strong> boundary layers adjacent to an aircraft<br />

w<strong>in</strong>g. The transition from lam<strong>in</strong>ar to turbulent flow is illustrated by water flow over the hull of a ship which<br />

<strong>in</strong>volves lam<strong>in</strong>ar flow at the bow followed by turbulent flow beh<strong>in</strong>d the bow wave and at the stern of the<br />

ship. The broad extent of the white foam of seawater along the side and the stern of a ship illustrates the<br />

considerable energy dissipation produced by the turbulence. The boundary layer of a stalled aircraft w<strong>in</strong>g<br />

is another example. At a high angle of attack, the airflow on the lower surface of the w<strong>in</strong>g rema<strong>in</strong>s lam<strong>in</strong>ar,<br />

that is, the stream velocity profile, relative to the w<strong>in</strong>g, <strong>in</strong>creases smoothly from zero at the w<strong>in</strong>g surface<br />

outwards until it meets the ambient air velocity on the outer surface of the boundary layer which is the order<br />

of a millimeter thick. The flow on the top surface of the w<strong>in</strong>g <strong>in</strong>itially is lam<strong>in</strong>ar before becom<strong>in</strong>g turbulent<br />

at which po<strong>in</strong>t the boundary layer rapidly <strong>in</strong>creases <strong>in</strong> thickness. Further back the airflow detaches from<br />

the w<strong>in</strong>g surface and large-scale vortex structures lead to a wide boundary layer comparable <strong>in</strong> thickness to<br />

the chord of the w<strong>in</strong>g with vortex motion that leads to the airflow revers<strong>in</strong>g its direction adjacent to the<br />

upper surface of the w<strong>in</strong>g which greatly <strong>in</strong>creases drag. When the vortices beg<strong>in</strong> to shed off the bounded<br />

surface they do so at a certa<strong>in</strong> frequency which can cause vibrations that can lead to structural failure if the<br />

frequency of the shedd<strong>in</strong>g vortices is close to the resonance frequency of the structure.<br />

Considerable time and effort are expended by aerodynamicists and hydrodynamicists design<strong>in</strong>g aircraft<br />

w<strong>in</strong>gs and ship hulls to maximize the length of lam<strong>in</strong>ar region of the boundary layer to m<strong>in</strong>imize drag.<br />

When the Reynolds number is large the slightest imperfections <strong>in</strong> the shape of w<strong>in</strong>g, such as a speck of<br />

dust, can trigger the transition from lam<strong>in</strong>ar to turbulent flow. The boundaries between adjacent large-scale<br />

coherent structures are sensitively identified <strong>in</strong> computer simulations by large divergence of the streaml<strong>in</strong>es<br />

at any separatrix. A large positive, f<strong>in</strong>ite-time, Lyapunov exponent identifies divergence of the streaml<strong>in</strong>es<br />

which occurs at a separatrix between adjacent large-scale coherent vortex structures, whereas the Lyapunov<br />

exponents are negative for converg<strong>in</strong>g streaml<strong>in</strong>es with<strong>in</strong> any coherent structure. Computations of turbulent<br />

flow often comb<strong>in</strong>e the use of f<strong>in</strong>ite-time Lyapunov exponents to identify coherent structures, plus Lagrangian<br />

mechanics for the equations of motion s<strong>in</strong>ce the Lagrangian is a scalar function, it is frame <strong>in</strong>dependent, and<br />

it gives far better results for fluid motion than us<strong>in</strong>g Newtonian mechanics. Thus the Lagrangian approach <strong>in</strong><br />

the cont<strong>in</strong>ua is used extensively for calculations <strong>in</strong> aerodynamics, hydrodynamics, and studies of atmospheric<br />

phenomena such as convection, hurricanes, tornadoes, etc.


16.9. SUMMARY AND IMPLICATIONS 455<br />

16.9 Summary and implications<br />

The goal of this chapter is to provide a glimpse <strong>in</strong>to the classical mechanics of the cont<strong>in</strong>ua which <strong>in</strong>troduces<br />

the Lagrangian density and Hamiltonian density formulations of classical mechanics.<br />

Lagrangian density formulation: In three dimensional Lagrangian density L(q q<br />

<br />

∇ · q) is<br />

related to the Lagrangian by tak<strong>in</strong>g the volume <strong>in</strong>tegral of the Lagrangian density.<br />

Z<br />

= L(q q<br />

∇ · q)<br />

(1621)<br />

Apply<strong>in</strong>g Hamilton’s Pr<strong>in</strong>ciple to the three-dimensional Lagrangian density leads to the follow<strong>in</strong>g set of<br />

differential equations of motion<br />

à ! à ! à ! à !<br />

L<br />

+ L<br />

+ L<br />

+ L<br />

− L<br />

<br />

q<br />

<br />

q<br />

<br />

q<br />

<br />

q<br />

q =0<br />

(1622)<br />

<br />

<br />

<br />

<br />

Hamiltonian density formulation: In the limit that the coord<strong>in</strong>ates are cont<strong>in</strong>uous, then the Hamiltonian<br />

density can be expressed <strong>in</strong> terms of a volume <strong>in</strong>tegral over the momentum density and the Lagrangian<br />

density L where<br />

π ≡ L<br />

(1627)<br />

˙q<br />

Then the obvious def<strong>in</strong>ition of the Hamiltonian density H is<br />

Z Z<br />

= H = (π · ˙q−L) (1628)<br />

where the Hamiltonian density is given by<br />

H =π · ˙q−L<br />

(1629)<br />

These Lagrangian and Hamiltonian density formulations are of considerable importance to field theory<br />

and fluid mechanics.<br />

L<strong>in</strong>ear elastic solids: The theory of cont<strong>in</strong>uous systems was applied to the case of l<strong>in</strong>ear elastic solids.<br />

The stress tensor T is a rank 2 tensor def<strong>in</strong>ed as the ratio of the force vector F and the surface element<br />

vector A. That is, the force vector is given by the <strong>in</strong>ner product of the stress tensor T and the surface<br />

element vector A.<br />

F = T·A<br />

(1633)<br />

The stra<strong>in</strong> tensor σ also is a rank 2 tensor def<strong>in</strong>ed as the ratio of the stra<strong>in</strong> vector ξ and <strong>in</strong>f<strong>in</strong>itessimal<br />

area A<br />

ξ = σ·A<br />

(1638)<br />

where the component form of the rank 2 stra<strong>in</strong> tensor is<br />

1 1 1<br />

σ = 1 1 2 3<br />

2 2 2<br />

2<br />

1 2 3<br />

(1639)<br />

¯<br />

¯¯¯¯¯¯¯<br />

3 3 3<br />

1 2 3<br />

The modulus of elasticity is def<strong>in</strong>ed as the slope of the stress-stra<strong>in</strong> curve. For l<strong>in</strong>ear, homogeneous,<br />

elastic matter, the potential energy density separates <strong>in</strong>to diagonal and off-diagonal components of the<br />

stra<strong>in</strong> tensor<br />

"<br />

= 1 X ( ) 2 +2 X #<br />

( ) 2 (1642)<br />

2<br />

<br />

<br />

where the constants and are Lamé’s moduli of elasticity which are positive. The stress tensor is related<br />

to the stra<strong>in</strong> tensor by<br />

X<br />

µ<br />

<br />

= <br />

+ + <br />

X<br />

<br />

= +2 <br />

(1643)


456 CHAPTER 16. ANALYTICAL FORMULATIONS FOR CONTINUOUS SYSTEMS<br />

Electromagnetic field theory: The rank 2 Maxwell stress tensor T has components<br />

µ<br />

≡ 0 − 1 <br />

2 2 + 1 µ<br />

− 1 <br />

0 2 2<br />

(1671)<br />

The divergence theorem allows the total electromagnetic force, act<strong>in</strong>g of the volume to be written as<br />

Z µ<br />

I<br />

Z<br />

S<br />

<br />

F= ∇ · T− 0 0 = T·a− 0 <br />

<br />

0 Sdτ<br />

(1674)<br />

<br />

The total momentum flux density is given by<br />

<br />

(π + π )=∇ · T<br />

(1679)<br />

where the electromagnetic field momentum density is given by the Poynt<strong>in</strong>g vector S as π = 0 0 S.<br />

Ideal fluid dynamics:<br />

Mass conservation leads to the cont<strong>in</strong>uity equation<br />

Euler’s hydrodynamic equation gives<br />

<br />

+ ∇· (v) =0<br />

(1682)<br />

v<br />

+(v · ∇) v = − 1 ∇ ( + )<br />

(1690)<br />

where is the scalar gravitational potential. If the flow is irrotational and time <strong>in</strong>dependent then<br />

µ 1<br />

2 2 + + <br />

<br />

= constant (1694)<br />

Viscous fluid dynamics: For <strong>in</strong>compressible flow the stress tensor term simplifies to ∇ · T =∇ 2 v.Then<br />

the Navier-Stokes equation becomes<br />

∙ ∇v¸<br />

v<br />

<br />

+ v · = −∇ + ∇ 2 v+f (1698)<br />

where ∇ 2 v is the viscosity drag term. The left-hand side of equation 1698 represents the rate of change<br />

of momentum per unit volume while the right-hand side represents the summation of the forces per unit<br />

volume that are act<strong>in</strong>g.<br />

The Reynolds number is a dimensionless number that characterizes the ratio of <strong>in</strong>ertial forces to viscous<br />

forces <strong>in</strong> a viscous medium. The evolution of flow from lam<strong>in</strong>ar flow to turbulent flow, with <strong>in</strong>crease of<br />

Reynolds number, was discussed.<br />

The classical mechanics of cont<strong>in</strong>uous fields encompasses a remarkably broad range of phenomena with<br />

important applications to lam<strong>in</strong>ar and turbulent fluid flow, gravitation, electromagnetism, relativity, and<br />

quantum fields.


Chapter 17<br />

Relativistic mechanics<br />

17.1 Introduction<br />

Newtonian mechanics <strong>in</strong>corporates the Newtonian concept of the complete separation of space and time.<br />

This theory reigned supreme from <strong>in</strong>ception, <strong>in</strong> 1687, until November 1905 when E<strong>in</strong>ste<strong>in</strong> pioneered the<br />

Special Theory of Relativity. Relativistic mechanics underm<strong>in</strong>es the Newtonian concepts of absoluteness of<br />

time that is <strong>in</strong>herent to Newton’s formulation, as well as when recast <strong>in</strong> the Lagrangian and Hamiltonian<br />

formulations of classical mechanics. Relativistic mechanics has had a profound impact on twentieth-century<br />

physics and the philosophy of science. <strong>Classical</strong> mechanics is an approximation of relativistic mechanics<br />

that is valid for velocities much less than the velocity of light <strong>in</strong> vacuum. The term “relativity” refers to<br />

the fact that physical measurements are always made relative to some chosen reference frame. Naively one<br />

may th<strong>in</strong>k that the transformation between different reference frames is trivial and conta<strong>in</strong>s little underly<strong>in</strong>g<br />

physics. However, E<strong>in</strong>ste<strong>in</strong> showed that the results of measurements depend on the choice of coord<strong>in</strong>ate<br />

system, which revolutionized our concept of space and time.<br />

E<strong>in</strong>ste<strong>in</strong>’s work on relativistic mechanics comprised two major advances. The first advance is the 1905<br />

Special Theory of Relativity which refers to nonaccelerat<strong>in</strong>g frames of reference. The second major advance<br />

was the 1916 General Theory of Relativity which considers accelerat<strong>in</strong>g frames of reference and their relation<br />

to gravity. The Special Theory is a limit<strong>in</strong>g case of the General Theory of Relativity. The mathematically<br />

complex General Theory of Relativity is required for describ<strong>in</strong>g accelerat<strong>in</strong>g frames, gravity, plus related<br />

topics like Black Holes, or extremely accurate time measurements <strong>in</strong>herent to the Global Position<strong>in</strong>g System.<br />

The present discussion will focus primarily on the mathematically simple Special Theory of Relativity s<strong>in</strong>ce it<br />

encompasses most of the physics encountered <strong>in</strong> atomic, nuclear and high energy physics. This chapter uses<br />

the basic concepts of the Special Theory of Relativity to <strong>in</strong>vestigate the implications of extend<strong>in</strong>g Newtonian,<br />

Lagrangian and Hamiltonian formulations of classical mechanics <strong>in</strong>to the relativistic doma<strong>in</strong>. The Lorentz<strong>in</strong>variant<br />

extended Hamiltonian and Lagrangian formalisms are <strong>in</strong>troduced s<strong>in</strong>ce they are applicable to the<br />

Special Theory of Relativity. The General Theory of Relativity <strong>in</strong>corporates the gravitational force as a<br />

geodesic phenomena <strong>in</strong> a four-dimensional Reimannian structure based on space, time, and matter. A<br />

superficial outl<strong>in</strong>e is given to the fundamental concepts and evidence that underlie the General Theory of<br />

Relativity.<br />

17.2 Galilean Invariance<br />

As discussed <strong>in</strong> chapter 23, an <strong>in</strong>ertial frame is one <strong>in</strong> which Newton’s Laws of motion apply. Inertial frames<br />

are non-accelerat<strong>in</strong>g frames so that pseudo forces are not <strong>in</strong>duced. All reference frames mov<strong>in</strong>g at constant<br />

velocity relative to an <strong>in</strong>ertial reference, are <strong>in</strong>ertial frames. Newton’s Laws of nature are the same <strong>in</strong> all<br />

<strong>in</strong>ertial frames of reference and therefore there is no way of determ<strong>in</strong><strong>in</strong>g absolute motion because no <strong>in</strong>ertial<br />

frame is preferred over any other. This is called Galilean-Newtonian <strong>in</strong>variance. Galilean <strong>in</strong>variance assumes<br />

that the concepts of space and time are completely separable. Time is assumed to be an absolute quantity<br />

that is <strong>in</strong>variant to transformations between coord<strong>in</strong>ate systems <strong>in</strong> relative motion. Also the element of<br />

length is the same <strong>in</strong> different Galilean frames of reference.<br />

457


458 CHAPTER 17. RELATIVISTIC MECHANICS<br />

Consider two coord<strong>in</strong>ate systems shown <strong>in</strong> figure 171, where the primed frame is mov<strong>in</strong>g along the <br />

axis of the fixed unprimed frame. A Galilean transformation implies that the follow<strong>in</strong>g relations apply;<br />

0 1 = 1 − (17.1)<br />

0 2 = 2<br />

0 3 = 3<br />

0 = <br />

Note that at any <strong>in</strong>stant the <strong>in</strong>f<strong>in</strong>itessimal units of length<br />

<strong>in</strong> the two systems are identical s<strong>in</strong>ce<br />

2 =<br />

3X<br />

2 =<br />

=1<br />

3X<br />

=1<br />

02<br />

= 02 (17.2)<br />

These are the mathematical expression of the Newtonian idea<br />

of space and time. An immediate consequence of the Galilean<br />

transformation is that the velocity of light must differ <strong>in</strong> different<br />

<strong>in</strong>ertial reference frames.<br />

At the end of the 19 century physicists thought they had<br />

discovered a way of identify<strong>in</strong>g an absolute <strong>in</strong>ertial frame of<br />

reference, that is, it must be the frame of the medium that<br />

transmits light <strong>in</strong> vacuum. Maxwell’s laws of electromagnetism<br />

predict that electromagnetic radiation <strong>in</strong> vacuum travels at =<br />

1<br />

√ <br />

=2998 × 10 8 . Maxwell did not address <strong>in</strong> what<br />

frame of reference that this speed applied. In the n<strong>in</strong>eteenth<br />

x2 x’ 2<br />

x<br />

x’<br />

3 3<br />

x1 x’ 1<br />

Figure 17.1: Motion of the primed frame<br />

along the 1 axis with velocity relative to<br />

the parallel unprimed frame.<br />

century all wave phenomena were transmitted by some medium, such as waves on a str<strong>in</strong>g, water waves,<br />

sound waves <strong>in</strong> air. Physicists thus envisioned that light was transmitted by some unobserved medium which<br />

they called the ether. This ether had mystical properties, it existed everywhere, even <strong>in</strong> outer space, and yet<br />

had no other observed consequences. The ether obviously should be the absolute frame of reference.<br />

In the 1880 0 , Michelson and Morley performed an experiment<br />

<strong>in</strong> Cleveland to try to detect this ether. They transmitted<br />

light back and forth along two perpendicular paths <strong>in</strong> an <strong>in</strong>terferometer,<br />

shown <strong>in</strong> figure 172, and assumed that the earth’s<br />

motion about the sun led to movement through the ether.<br />

The time taken to travel a return trip takes longer <strong>in</strong> a<br />

mov<strong>in</strong>g medium, if the medium moves <strong>in</strong> the direction of the<br />

motion, compared to travel <strong>in</strong> a stationary medium. For example,<br />

you lose more time mov<strong>in</strong>g aga<strong>in</strong>st a headw<strong>in</strong>d than<br />

you ga<strong>in</strong> travell<strong>in</strong>g back with the w<strong>in</strong>d. The time difference<br />

∆ for a round trip to a distance , between travell<strong>in</strong>g <strong>in</strong> the<br />

direction of motion <strong>in</strong> the ether, versus travell<strong>in</strong>g the same distance<br />

perpendicular to the movement <strong>in</strong> the ether, is given by<br />

¡<br />

∆ ≈ <br />

¢ 2<br />

where is the relative velocity of the ether and <br />

is the velocity of light.<br />

Interference fr<strong>in</strong>ges between perpendicular light beams <strong>in</strong><br />

an optical <strong>in</strong>terferometer provides an extremely sensitive measure<br />

of this time difference. Michelson and Morley observed no<br />

measurable time difference at any time dur<strong>in</strong>g the year, that<br />

is, the relative motion of the earth with<strong>in</strong> the ether is less than<br />

Light<br />

source<br />

L<br />

C<br />

A<br />

Mirror<br />

Semi-transparent<br />

mirror<br />

L<br />

Mirror<br />

Figure 17.2: The Michelson <strong>in</strong>terferometer<br />

used for the Michelson-Morley experiment.<br />

Interference of the two beams of coherent<br />

light leads to fr<strong>in</strong>ges that depends on the<br />

differences <strong>in</strong> phase along the two paths.<br />

16 the velocity of the earth around the sun. Their conclusion was either, that the ether was dragged along<br />

with the earth, or the velocity of light was dependent on the velocity of the source, but these did not jibe<br />

with other observations. Their disappo<strong>in</strong>tment at the failure of this experiment to detect evidence for an absolute<br />

<strong>in</strong>ertial frame is important and confounded physicists for two decades until E<strong>in</strong>ste<strong>in</strong>’s Special Theory<br />

of Relativity expla<strong>in</strong>ed the result.<br />

B<br />

v


17.3. SPECIAL THEORY OF RELATIVITY 459<br />

17.3 Special Theory of Relativity<br />

17.3.1 E<strong>in</strong>ste<strong>in</strong> Postulates<br />

In November 1905, at the age of 26, E<strong>in</strong>ste<strong>in</strong> published a sem<strong>in</strong>al paper entitled ”On the electrodynamics of<br />

mov<strong>in</strong>g bodies”. He considered the relation between space and time <strong>in</strong> <strong>in</strong>ertial frames of reference that are<br />

<strong>in</strong> relative motion. In this paper he made the follow<strong>in</strong>g postulates.<br />

1) The laws of nature are the same <strong>in</strong> all <strong>in</strong>ertial frames of reference.<br />

2) The velocity of light <strong>in</strong> vacuum is the same <strong>in</strong> all <strong>in</strong>ertial frames of reference.<br />

Note that E<strong>in</strong>ste<strong>in</strong>’s first postulate, coupled with Maxwell’s equations, leads to the statement that the<br />

velocity of light <strong>in</strong> vacuum is a universal constant. Thus the second postulate is unnecessary s<strong>in</strong>ce it is an<br />

obvious consequence of the first postulate plus Maxwell’s equations which are basic laws of physics. This<br />

second postulate expla<strong>in</strong>ed the null result of the Michelson-Morley experiment. However, it was not this<br />

experimental result that led E<strong>in</strong>ste<strong>in</strong> to the theory of special relativity; he deduced the Special Theory of<br />

Relativity from consideration of Maxwell’s equations of electromagnetism. Although E<strong>in</strong>ste<strong>in</strong>’s postulates<br />

appear reasonable, they lead to the follow<strong>in</strong>g surpris<strong>in</strong>g implications.<br />

17.3.2 Lorentz transformation<br />

Galilean <strong>in</strong>variance leads to violation of the E<strong>in</strong>ste<strong>in</strong> postulate that the velocity of light is a universal constant<br />

<strong>in</strong> all frames of reference. It is necessary to assume a new transformation law that renders physical<br />

laws relativistically <strong>in</strong>variant. Maxwell’s equations are relativistically <strong>in</strong>variant, which led to some electromagnetic<br />

phenomena that could not be expla<strong>in</strong>ed us<strong>in</strong>g Galilean <strong>in</strong>variance. In 1904 Lorentz proposed a new<br />

transformation to replace the Galilean transformation <strong>in</strong> order to expla<strong>in</strong> such electromagnetic phenomena.<br />

E<strong>in</strong>ste<strong>in</strong>’s genius was that he derived the transformation, that had been proposed by Lorentz, directly from<br />

the postulates of the Special Theory of Relativity. The Lorentz transformation satisfies E<strong>in</strong>ste<strong>in</strong>’s theory of<br />

relativity, and has been confirmed to be correct by many experiments.<br />

For the geometry shown <strong>in</strong> figure 171, theLorentz transformations are:<br />

where the Lorentz factor<br />

The <strong>in</strong>verse transformations are<br />

= ( 0 + 0 ) (17.5)<br />

= 0<br />

= 0<br />

<br />

= <br />

µ 0 + 0<br />

2<br />

The Lorentz factor, def<strong>in</strong>ed above, is the key feature<br />

differentiat<strong>in</strong>g the Lorentz transformations from the Galilean<br />

transformation. Note that ≥ 1; also → 10 as → 0 and<br />

<strong>in</strong>creases to <strong>in</strong>f<strong>in</strong>ity as <br />

→ 1 as illustrated <strong>in</strong> figure 173. A<br />

useful fact that will be used later is that for <br />

1;<br />

0 = ( − ) (17.3)<br />

0 = <br />

0 = ³<br />

0 = − ´<br />

2<br />

1<br />

≡ q<br />

1 − ¡ ¢<br />

(17.4)<br />

2<br />

<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0<br />

v<br />

c<br />

→ 1+ 1 ³ <br />

´2<br />

2 <br />

Limit for <br />

Note that for then =1and the Lorentz transformation<br />

is identical to the Galilean transformation.<br />

Figure 17.3: The dependence of the Lorentz<br />

factor on


460 CHAPTER 17. RELATIVISTIC MECHANICS<br />

Tick!<br />

Tick!<br />

d<br />

d<br />

Tock! Tock! Tock!<br />

a. b.<br />

Figure 17.4: The observer and mirror are at rest <strong>in</strong> the left-hand frame (a). The light beam takes a time<br />

∆ = <br />

to travel to the mirror. In the right-hand frame (b) the source and mirror are travell<strong>in</strong>g at a velocity<br />

relative to the observer. The light travels further <strong>in</strong> the right-hand frame of reference (b) than is the<br />

stationary frame (a). S<strong>in</strong>ce E<strong>in</strong>ste<strong>in</strong> states that the velocity of light is the same <strong>in</strong> both frames of reference<br />

then the time <strong>in</strong>terval must by larger <strong>in</strong> frame (b) s<strong>in</strong>ce the light travels further than <strong>in</strong> (a).<br />

17.3.3 Time Dilation:<br />

Consider that a clock is fixed at 0 <strong>in</strong> a mov<strong>in</strong>g frame and measures the time <strong>in</strong>terval between two events<br />

<strong>in</strong> the mov<strong>in</strong>g frame, i.e. ∆ 0 = 0 1 − 0 2. Accord<strong>in</strong>g to the Lorentz transformation, the times <strong>in</strong> the fixed<br />

frame are given by:<br />

µ <br />

1 = 0 1 + 0 0<br />

2 (17.6)<br />

µ <br />

2 = 0 2 + 0 0<br />

2<br />

Thus the time <strong>in</strong>terval is given by:<br />

2 − 1 = ( 0 2 − 0 1) (17.7)<br />

The time between events <strong>in</strong> the rest frame of the clock, ∆ ≡ ∆ 0 is called the proper time which always<br />

is the shortest time measured for a given event and is represented by the symbol . Thatis<br />

∆ = ∆ 0 = ∆ (17.8)<br />

Note that the time <strong>in</strong>terval for any other frame of reference, mov<strong>in</strong>g with respect to the clock frame, will<br />

show larger time <strong>in</strong>tervals because ≥ 10 which implies that the fixed frame perceives that the mov<strong>in</strong>g<br />

clockisslowbythefactor.<br />

The plausibility of this time dilation can be understood by look<strong>in</strong>g at the simple geometry of the space<br />

ship example shown <strong>in</strong> Figure 174. Pretend that the clock <strong>in</strong> the proper frame of the space ship is based on<br />

thetimeforthelighttotraveltoandfromthemirror<strong>in</strong>thespaceship. Inthisproperframethelighthas<br />

the shortest distance to travel, and the proper transit time is<br />

∆ = 2 <br />

(17.9)<br />

In the fixed frame, the component of velocity <strong>in</strong> the direction of the mirror is √ 2 − 2 us<strong>in</strong>g the Pythagorus<br />

theorem, assum<strong>in</strong>g that the light cannot travel faster the . Thus the transit time towards and back from<br />

the mirror must be<br />

2<br />

∆ = q<br />

1 − ¡ ¢<br />

= ∆ (17.10)<br />

2<br />

<br />

which is the predicted time dilation.


17.3. SPECIAL THEORY OF RELATIVITY 461<br />

There are many experimental verifications of time dilation <strong>in</strong> physics. For example, a stationary muon<br />

has a mean lifetime of =2 sec, whereas the lifetime of a fast mov<strong>in</strong>g muon, produced <strong>in</strong> the upper<br />

atmosphere by high-energy cosmic rays, was observed <strong>in</strong> 1941 to be longer and given by as described <strong>in</strong><br />

example 171. In1972 Hafely and Keat<strong>in</strong>g used four accurate cesium atomic clocks to confirm time dilation.<br />

Two clocks were flown on regularly scheduled airl<strong>in</strong>es travell<strong>in</strong>g around the World, one westward and the<br />

other eastward. The other two clocks were used for reference. The westward mov<strong>in</strong>g clock was slow by<br />

(273 ± 7) compared to the predicted value of (275 ± 10) sec. The Global Position<strong>in</strong>g System of 24<br />

geosynchronous satellites is used for locat<strong>in</strong>g positions to with<strong>in</strong> a few meters. It has an accuracy of a few<br />

nanoseconds which requires allowance for time dilation and is a daily tribute to the correctness of E<strong>in</strong>ste<strong>in</strong>’s<br />

Theory of Relativity.<br />

17.3.4 Length Contraction<br />

The Lorentz transformation leads to a contraction of the apparent length of an object <strong>in</strong> a mov<strong>in</strong>g frame<br />

as seen from a fixed frame. The length of a ruler <strong>in</strong> its own frame of reference is called the proper length.<br />

Consider an accurately measured rod of known proper length = 0 2 − 0 1 thatis,atrest<strong>in</strong>themov<strong>in</strong>g<br />

primed frame. The locations of both ends of this rod are measured at a given time <strong>in</strong> the stationary frame,<br />

1 = 2 by tak<strong>in</strong>g a photograph of the mov<strong>in</strong>g rod. The correspond<strong>in</strong>g locations <strong>in</strong> the mov<strong>in</strong>g frame are:<br />

S<strong>in</strong>ce 2 = 1 , the measured lengths <strong>in</strong> the two frames are related by:<br />

0 2 = ( 2 − 2 ) (17.11)<br />

0 1 = ( 1 − 1 )<br />

0 2 − 0 1 = ( 2 − 1 ) (17.12)<br />

That is, the lengths are related by:<br />

= 1 (17.13)<br />

Note that the mov<strong>in</strong>g rod appears shorter <strong>in</strong> the direction of motion. As → the apparent length<br />

shr<strong>in</strong>ks to zero <strong>in</strong> the direction of motion while the dimensions perpendicular to the direction of motion are<br />

unchanged. This is called the Lorentz contraction. If you could ride your bicycle at close to the speed of<br />

light, you would observe that stationary cars, build<strong>in</strong>gs, people, all would appear to be squeezed th<strong>in</strong> along<br />

the direction that you are travell<strong>in</strong>g. Also objects that are further away down any side street would be<br />

distorted <strong>in</strong> the direction of travel. A photograph taken by a stationary observer would show the mov<strong>in</strong>g<br />

bicycle to be Lorentz contracted along the direction of travel and the stationary objects would be normal.<br />

17.3.5 Simultaneity<br />

The Lorentz transformations imply a new philosophy of space and time. A surpris<strong>in</strong>g consequence is that<br />

the concept of simultaneity is frame dependent <strong>in</strong> contrast to the prediction of Newtonian mechanics.<br />

Consider that two events occur <strong>in</strong> frame at ( 1 1 ) and ( 2 2 ) In frame 0 these two events occur at<br />

( 0 1 0 1) and ( 0 2 0 2) From the Lorentz transformation the time difference is<br />

∙<br />

0 2 − 0 1 = ( 2 − 1 ) − ( ¸<br />

2 − 1 )<br />

2<br />

(17.14)<br />

If an event is simultaneous <strong>in</strong> frame that is ( 2 − 1 )=0then<br />

∙ ¸<br />

<br />

0 2 − 0 (1 − 2 )<br />

1 = <br />

2<br />

(17.15)<br />

Thus the event is not simultaneous <strong>in</strong> frame 0 if ( 2 − 1 )= 6=0 That is, an event that is simultaneous<br />

<strong>in</strong> one frame is not simultaneous <strong>in</strong> the other frame if the events are spatially separated. The equivalent<br />

statement is that for two clocks, spatially separated by a distance , which are synchronized <strong>in</strong> their rest<br />

frame, then <strong>in</strong> a mov<strong>in</strong>g frame they are not simultaneous.


462 CHAPTER 17. RELATIVISTIC MECHANICS<br />

v<br />

L<br />

2<br />

L<br />

2<br />

Figure 17.5: If lightn<strong>in</strong>g strikes the front and rear of the carriage simultaneously, accord<strong>in</strong>g to the man <strong>in</strong><br />

the fixed frame, then the woman <strong>in</strong> the mov<strong>in</strong>g frame sees the flash from the front first s<strong>in</strong>ce she is mov<strong>in</strong>g<br />

towards that approach<strong>in</strong>g wavefront dur<strong>in</strong>g the transit time of the light. Thus if the length of the carriage<br />

<strong>in</strong> the stationary frame is ( 2 − 1 )= thenthetimedifference is ∆ 0 = <br />

<br />

2 .<br />

E<strong>in</strong>ste<strong>in</strong> discussed the example shown <strong>in</strong> figure 175, where lightn<strong>in</strong>g strikes both ends of a tra<strong>in</strong> simultaneously<br />

<strong>in</strong> the stationary earth frame of reference. A woman on the tra<strong>in</strong> will see that the strikes are<br />

not simultaneous s<strong>in</strong>ce the wavefront from the front of the carriage will be seen first because she is mov<strong>in</strong>g<br />

forward dur<strong>in</strong>g the time the light from the two lightn<strong>in</strong>g flashes is travell<strong>in</strong>g towards her. As a consequence<br />

she observes that the two lightn<strong>in</strong>g flashes are not simultaneous. This expla<strong>in</strong>s why measurement of the<br />

length of a mov<strong>in</strong>g rod, performed by simultaneously locat<strong>in</strong>g both ends <strong>in</strong> the fixed frame, implies that the<br />

measurement occurs at different times for both ends <strong>in</strong> the mov<strong>in</strong>g frame result<strong>in</strong>g <strong>in</strong> a shorter apparent<br />

length. The lack of simultaneity expla<strong>in</strong>s why one can get the apparent <strong>in</strong>consistency that the mov<strong>in</strong>g bicyclist<br />

sees that the stationary street block to be length contracted, while <strong>in</strong> contrast, a pedestrian sees that<br />

the bicycle is length contracted.<br />

The concept of causality breaks down s<strong>in</strong>ce ( 0 2 − 0 1) can be either positive or negative, therefore the<br />

correspond<strong>in</strong>g ∆ can be positive of negative. A consequence of the lack of simultaneity is that the image<br />

shown by a photograph of a rapidly mov<strong>in</strong>g object is not a true representation of the mov<strong>in</strong>g object. Not<br />

only is the body contracted <strong>in</strong> the direction of travel, but also it appears distorted because light arriv<strong>in</strong>g from<br />

the far side of the body had to be emitted earlier, that is, when the body was at an earlier location, <strong>in</strong> order<br />

to reach the observer simultaneously with light from the near side. The relativistic snake paradox, addressed<br />

<strong>in</strong> Chapter 17 workshop exercise 1 is an example of the role of simultaneity <strong>in</strong> relativistic mechanics.<br />

17.1 Example: Muon lifetime<br />

Many people had trouble comprehend<strong>in</strong>g time dilation and Lorentz contraction predicted by the Special<br />

Theory of Relativity. The predictions appear to be crazy, but there are many examples where time dilation<br />

and Lorentz contraction are observed experimentally such as the decay <strong>in</strong> flight of the muon. At rest, the<br />

muon decays with a mean lifetime of 2 sec Muons are created high <strong>in</strong> the atmosphere due to cosmic ray<br />

bombardment. A typical muon travels at =0998 which corresponds to =15 Time dilation implies<br />

that the lifetime of the mov<strong>in</strong>g muon <strong>in</strong> the earth’s frame of reference is 30 . The speed of the muon is<br />

essentially <strong>in</strong> both frames of reference, and it would travel 600 <strong>in</strong> 2 and 9000 <strong>in</strong> 30 . In fact,<br />

it is observed that the muon does travel, on average, 9000 <strong>in</strong> the earth frame of reference before decay<strong>in</strong>g.<br />

Is this <strong>in</strong>consistent with the view of someone travell<strong>in</strong>g with the muon? In the muon’s mov<strong>in</strong>g frame, the<br />

lifetime is only 2 , but the Lorentz contraction of distance means that 9000 <strong>in</strong> the earth frame appears<br />

to be only 600 <strong>in</strong> the muon mov<strong>in</strong>g frame; a distance it travels is 2 sec. Thus<strong>in</strong>bothframesofreference<br />

we have consistent explanations, that is, the muon travels the height of the mounta<strong>in</strong> <strong>in</strong> one lifetime.


17.3. SPECIAL THEORY OF RELATIVITY 463<br />

17.2 Example: Relativistic Doppler Effect<br />

The relativistic Doppler effect is encountered frequently <strong>in</strong> physics and astronomy. Consider monochromatic<br />

electromagnetic radiation from a source, such as a star, that is mov<strong>in</strong>g towards the detector at a<br />

velocity . Dur<strong>in</strong>g the time ∆ <strong>in</strong>theframeofthereceiver,thesourceemits cycles of the s<strong>in</strong>usoidal<br />

waveform. Thus the length of this waveform, as seen by the receiver, is which equals<br />

The frequency as measured by the receiver is<br />

=( − )∆<br />

= =<br />

<br />

( − )∆<br />

Accord<strong>in</strong>g to the source, it emits waves of frequency 0 dur<strong>in</strong>g the proper time <strong>in</strong>terval ∆ 0 ,thatis<br />

= 0 ∆ 0<br />

This proper time <strong>in</strong>terval ∆ 0 , <strong>in</strong> the source frame, corresponds to a time <strong>in</strong>terval ∆ <strong>in</strong> the receiver frame<br />

where<br />

Thus the frequency measured by the receiver is<br />

=<br />

∆ = ∆ 0<br />

p<br />

1 0 1 − (<br />

<br />

(1 − ) = )2<br />

(1 − ) 0 =<br />

s<br />

1+<br />

1 − 0<br />

where ≡ <br />

. This formula for source and receiver approach<strong>in</strong>g each other also gives the correct answer for<br />

source and receiver reced<strong>in</strong>g if the sign of is changed.<br />

This relativistic Doppler Effect accounts for the red shift observed for light emitted by reced<strong>in</strong>g stars and<br />

galaxies, as well as many examples <strong>in</strong> atomic and nuclear physics <strong>in</strong>volv<strong>in</strong>g mov<strong>in</strong>g sources of electromagnetic<br />

radiation.<br />

17.3 Example: Tw<strong>in</strong> paradox<br />

A problem that troubled physicists for many years is called the tw<strong>in</strong> paradox. Consider two identical<br />

tw<strong>in</strong>s, Jack and Jill. Assume that Jill travels <strong>in</strong> a space ship at a speed of =4for 20 years, as measured<br />

by Jack’s clock, and then returns tak<strong>in</strong>g another 20 years, accord<strong>in</strong>g to Jack. Thus, Jack has aged 40 years<br />

by the time his tw<strong>in</strong> sister returns home. However, Jill’s clock measures 204 =5years for each half of the<br />

trip so that she th<strong>in</strong>ks she travelled for 10 years total time accord<strong>in</strong>g to her clock. Thus she has aged only 10<br />

years on the trip, that is, now she is 30 years younger that her tw<strong>in</strong> brother. Note that, accord<strong>in</strong>g to Jill, the<br />

distance she travelled out and back was 14 the distance accord<strong>in</strong>g to Jack, so she perceives no <strong>in</strong>consistency<br />

<strong>in</strong> her clock, and the speed of the space ship. This was called a paradox because some people claimed that<br />

Jill will perceive that the earth and Jack moved away at the same relative speed <strong>in</strong> the opposite direction and<br />

thus accord<strong>in</strong>g to Jill, Jack should be 30 years younger, not her. Moreover, some claimed that this problem<br />

issymmetricandthereforebothtw<strong>in</strong>smuststillbethesameages<strong>in</strong>cethereisnowayoftell<strong>in</strong>gwhowas<br />

mov<strong>in</strong>g away from whom. This argument is <strong>in</strong>correct because Jill was able to sense that she accelerated to<br />

=4which destroys the symmetry argument. The effectisobservedwithacceleratedbeamsofunstable<br />

nuclei such as the muon and was confirmed by the results of the experiment where cesium atomic clocks were<br />

flown around the Earth. Thus the Tw<strong>in</strong> paradox is not a paradox; the fact is that Jill will be younger than<br />

her tw<strong>in</strong> brother.


464 CHAPTER 17. RELATIVISTIC MECHANICS<br />

17.4 Relativistic k<strong>in</strong>ematics<br />

17.4.1 Velocity transformations<br />

Consider the two parallel coord<strong>in</strong>ate frames with the primed frame mov<strong>in</strong>g at a velocity along the 0 1 axis<br />

as shown <strong>in</strong> figure 171. Velocitiesofanobjectmeasured<strong>in</strong>bothframesaredef<strong>in</strong>ed to be<br />

= <br />

<br />

0 = 0 <br />

0<br />

(17.16)<br />

Us<strong>in</strong>g the Lorentz transformations 173 175 between the two frames mov<strong>in</strong>g with relative velocity along<br />

the 1 axis, gives that the velocity along the 0 1 axis is<br />

0 1 = 0 1<br />

0<br />

Similarly we get the velocities along the perpendicular 0 2 and 0 3 axes to be<br />

= 1 − <br />

− 2 1<br />

= 1 − <br />

1 − 1<br />

2 (17.17)<br />

0 2 = 0 2<br />

0 = 2<br />

1 − 1<br />

(17.18)<br />

2<br />

0 3 = 0 3<br />

0 = 3<br />

1 − 1<br />

2<br />

When 1<br />

<br />

→ 0 these velocity transformations become the usual Galilean relations for velocity addition.<br />

2<br />

Do not confuse u and u 0 with v; thatis,u and u 0 are the velocities of some object measured <strong>in</strong> the unprimed<br />

and primed frames of reference respectively, whereas v is the relative velocity of the orig<strong>in</strong> of one frame with<br />

respect to the orig<strong>in</strong> of the other frame.<br />

17.4.2 Momentum<br />

Us<strong>in</strong>g the classical def<strong>in</strong>ition of momentum, that is p =u, the l<strong>in</strong>ear momentum is not conserved us<strong>in</strong>g the<br />

above relativistic velocity transformations if the mass is a scalar quantity. This problem orig<strong>in</strong>ates from<br />

thefactthatbothx and have non-trivial transformations and thus u = x<br />

<br />

is frame dependent.<br />

L<strong>in</strong>ear momentum conservation can be reta<strong>in</strong>ed by redef<strong>in</strong><strong>in</strong>g momentum <strong>in</strong> a form that is identical <strong>in</strong><br />

all frames of reference, that is by referr<strong>in</strong>g to the proper time as measured <strong>in</strong> the rest frame of the mov<strong>in</strong>g<br />

object. Therefore we def<strong>in</strong>e relativistic l<strong>in</strong>ear momentum as<br />

p ≡ x<br />

= x <br />

(17.19)<br />

<br />

But we know the time dilation relation<br />

<br />

= q = (17.20)<br />

(1 − 2<br />

<br />

) 2<br />

Note that the <strong>in</strong> this relation refers to the velocity between the mov<strong>in</strong>g object and the frame; this is<br />

1<br />

quite different from the = which refers to the transformation between the two frames of reference.<br />

(1− 2<br />

2 )<br />

Thus the new relativistic def<strong>in</strong>ition of momentum is<br />

p ≡ x<br />

= x<br />

<br />

= u (17.21)<br />

The relativistic def<strong>in</strong>ition of l<strong>in</strong>ear momentum is the same as the classical def<strong>in</strong>ition with the rest mass<br />

replaced by the relativistic mass . 1<br />

1 Note that, until recently, the rest mass was denoted by 0 and the relativistic mass was referred to as . Modern texts<br />

denote the rest mass by and the relativistic mass by . This book follows the modern nomenclature for rest mass to avoid<br />

confusion.


17.4. RELATIVISTIC KINEMATICS 465<br />

17.4.3 Center of momentum coord<strong>in</strong>ate system<br />

The classical relations for handl<strong>in</strong>g the k<strong>in</strong>ematics of collid<strong>in</strong>g objects, carry over to special relativity when the<br />

relativistic def<strong>in</strong>ition of l<strong>in</strong>ear momentum, equation 1721, is assumed. That is, one can cont<strong>in</strong>ue to apply<br />

conservation of l<strong>in</strong>ear momentum. However, there is one important conceptual difference for relativistic<br />

dynamics <strong>in</strong> that the center of mass no longer is a mean<strong>in</strong>gful concept due to the <strong>in</strong>terrelation of mass<br />

and energy. However, this problem is elim<strong>in</strong>ated by consider<strong>in</strong>g the center of momentum coord<strong>in</strong>ate system<br />

which, as <strong>in</strong> the non-relativistic case, is the frame where the total l<strong>in</strong>ear momentum of the system is zero.<br />

Us<strong>in</strong>g the concept of center of momentum <strong>in</strong>corporates the formalism of classical non-relativistic k<strong>in</strong>ematics.<br />

17.4.4 Force<br />

Newton’s second law F = p<br />

<br />

is covariant under a Galilean transformation. In special relativity this def<strong>in</strong>ition<br />

also applies us<strong>in</strong>g the relativistic def<strong>in</strong>ition of momentum p. The fact that the relativistic momentum p<br />

is conserved <strong>in</strong> the force-free situation, leads naturally to us<strong>in</strong>g the def<strong>in</strong>ition of force to be<br />

F = p<br />

<br />

Then the relativistic momentum is conserved if F =0<br />

(17.22)<br />

17.4.5 Energy<br />

The classical def<strong>in</strong>ition of work done is def<strong>in</strong>ed by<br />

12 =<br />

Z 2<br />

1<br />

F·r = 2 − 1 (17.23)<br />

Assume 1 =0 let r = u and <strong>in</strong>sert the relativistic force relation <strong>in</strong> equation 1723, gives<br />

= =<br />

Z <br />

0<br />

<br />

( u) ·u = <br />

Integrate by parts, followed by algebraic manipulation, gives<br />

Z <br />

= 2 <br />

− q<br />

0 1 − 2<br />

2<br />

=<br />

2<br />

q + q<br />

2<br />

1 − 2<br />

2 1 − 2<br />

2 µ<br />

Z <br />

= 2 + 2 r<br />

0<br />

( ) (17.24)<br />

1 − 2<br />

2 − 2<br />

<br />

1 − 2<br />

2 − 2 = 2 ( − 1) (17.25)<br />

Def<strong>in</strong>e the rest energy 0<br />

0 ≡ 2 (17.26)<br />

and total relativistic energy <br />

then equation 1725 can be written as<br />

≡ 2 (17.27)<br />

= + 0 = 2 (17.28)<br />

This is the famous E<strong>in</strong>ste<strong>in</strong> relativistic energy that relates the equivalence of mass and energy. The total<br />

relativistic energy is a conserved quantity <strong>in</strong> nature. It is an extension of the conservation of energy and<br />

manifestations of the equivalence of energy and mass occur extensively <strong>in</strong> the real world.<br />

In nuclear physics we often convert mass to energy and back aga<strong>in</strong> to mass. For example, gamma<br />

rays with energies greater than 1022, which are pure electromagnetic energy, can be converted to an<br />

electron plus positron both of which have rest mass. The positron can then annihilate a different electron <strong>in</strong><br />

another atom result<strong>in</strong>g <strong>in</strong> emission of two 511 gamma rays <strong>in</strong> back to back directions to conserve l<strong>in</strong>ear<br />

momentum. A dramatic example of E<strong>in</strong>ste<strong>in</strong>’s equation is a nuclear reactor. One gram of material, the mass


466 CHAPTER 17. RELATIVISTIC MECHANICS<br />

of a paper clip, provides =9× 10 13 joules. This is the daily output of a 1 nuclear power station or<br />

the explosive power of the Nagasaki or Hiroshima bombs.<br />

As the velocity of a particle approaches then and the relativistic mass both approach <strong>in</strong>f<strong>in</strong>ity.<br />

This means that the force needed to accelerate the mass also approaches <strong>in</strong>f<strong>in</strong>ity, and thus no particle can<br />

exceed the velocity of light. The energy cont<strong>in</strong>ues to <strong>in</strong>crease not by <strong>in</strong>creas<strong>in</strong>g the velocity but by <strong>in</strong>crease<br />

of the relativistic mass. Although the relativistic relation for k<strong>in</strong>etic energy is quite different from the<br />

Newtonian relation, the Newtonian form is obta<strong>in</strong>ed for the case of <strong>in</strong> that<br />

= 2 (1 − 2<br />

2 )− 1 2 − 2 = 2 (1 + 1 2<br />

2 2 + ···) − 2 = 1 2 2 (17.29)<br />

An especially useful relativistic relation that can be derived from the above is<br />

2 = 2 2 + 2 0 (17.30)<br />

This is useful because it provides a simple relation between total energy of a particle and its relativistic<br />

l<strong>in</strong>ear momentum plus rest energy.<br />

17.4 Example: Rocket propulsion<br />

Consider a rocket, hav<strong>in</strong>g <strong>in</strong>itial mass is accelerated <strong>in</strong> a straight l<strong>in</strong>e <strong>in</strong> free space by exhaust<strong>in</strong>g<br />

propellant at a constant speed relative to the rocket. Let bethespeedoftherocketrelativetoit’s<strong>in</strong>itial<br />

rest frame when its rest mass has decreased to . At this <strong>in</strong>stant the rocket is at rest <strong>in</strong> the <strong>in</strong>ertial frame<br />

0 . At a proper time + the rest mass is − and it has acquired a velocity <strong>in</strong>crement relative to<br />

0 and propellant of rest mass has been expelled with velocity relative to 0 . At proper time <strong>in</strong> 0<br />

the rest mass is 2 . At the time + energy conservation requires that<br />

0 ( − ) 2 + <br />

2 = 2<br />

At the same <strong>in</strong>stant, conservation of l<strong>in</strong>ear momentum requires<br />

0 ( − ) 0 − =0<br />

To first order these two equations simplify to<br />

=<br />

r<br />

³ <br />

´2<br />

1 −<br />

<br />

0 = <br />

<br />

Therefore<br />

0 = <br />

The velocity <strong>in</strong>crement 0 <strong>in</strong> frame 0 can be transformed back to frame us<strong>in</strong>g equation 175, that is<br />

+ = + µ ³ 0<br />

<br />

´2<br />

≈ + 1 − 0<br />

1+ 0<br />

<br />

2<br />

Equations and yieldadifferential equation for () of<br />

<br />

1 − ¡ <br />

¢<br />

2<br />

= <br />

<br />

<br />

Integrate the left-hand side between 0 and and the right-hand side between and gives<br />

µ<br />

1 1+<br />

<br />

³<br />

2 ln <br />

<br />

´<br />

1 − = − ln<br />

<br />

<br />

This reduces to<br />

<br />

= 1 − ¡ <br />

<br />

1+ ¡ <br />

<br />

¢ 2<br />

¢ 2<br />

When <br />

→ 0 this equation reduces to the non-relativistic answer given <strong>in</strong> equation 2123.<br />

()<br />

()


17.5. GEOMETRY OF SPACE-TIME 467<br />

17.5 Geometry of space-time<br />

17.5.1 Four-dimensional space-time<br />

In 1906 Po<strong>in</strong>caré showed that the Lorentz transformation can be regarded as a rotation <strong>in</strong> a 4-dimensional<br />

Euclidean space-time <strong>in</strong>troduced by add<strong>in</strong>g an imag<strong>in</strong>ary fourth space-time coord<strong>in</strong>ate to the three<br />

real spatial coord<strong>in</strong>ates. In 1908 M<strong>in</strong>kowski reformulated E<strong>in</strong>ste<strong>in</strong>’s Special Theory of Relativity <strong>in</strong> this 4-<br />

dimensional Euclidean space-time vector space and concluded that the spatial variables where ( =1 2 3) <br />

plus the time 0 = are equivalent variables and should be treated equally us<strong>in</strong>g a covariant representation<br />

of both space and time. The idea of us<strong>in</strong>g an imag<strong>in</strong>ary time axis to make space-time Euclidean was<br />

elegant, but it obscured the non-Euclidean nature of space-time as well as caus<strong>in</strong>g difficulties when generalized<br />

to non-<strong>in</strong>ertial accelerat<strong>in</strong>g frames <strong>in</strong> the General Theory of Relativity. As a consequence, the use of the<br />

imag<strong>in</strong>ary has been abandoned <strong>in</strong> modern work. M<strong>in</strong>kowski developed an alternative non-Euclidean<br />

metric that treats all four coord<strong>in</strong>ates () as a four-dimensional M<strong>in</strong>kowski metric with all coord<strong>in</strong>ates<br />

be<strong>in</strong>g real, and <strong>in</strong>troduces the required m<strong>in</strong>us sign explicitly.<br />

Analogous to the usual 3-dimensional cartesian coord<strong>in</strong>ates, the displacement four vector s is def<strong>in</strong>ed<br />

us<strong>in</strong>g the four components along the four unit vectors <strong>in</strong> either the unprimed or primed coord<strong>in</strong>ate frames.<br />

s = 0 ê 0 + 1 ê 1 + 2 ê 2 + 3 ê 3 = 00 ê 0 0 + 01 ê 0 1 + 02 ê 0 2 + 03 ê 0 3 (17.31)<br />

The convention used is that greek subscripts (covariant) or superscripts (contravariant) designate a four<br />

vector with 0 ≤ ≤ 3 The covariant unit vectors ê are written with the subscript which has 4 values<br />

0 ≤ ≤ 3. As described <strong>in</strong> appendix 3, us<strong>in</strong>g the E<strong>in</strong>ste<strong>in</strong> convention the components are written with<br />

the contravariant superscript where the time axis 0 = , while the spatial coord<strong>in</strong>ates, expressed <strong>in</strong><br />

cartesian coord<strong>in</strong>ates, are 1 = , 2 = , and 3 = . With respect to a different (primed) unit vector basis<br />

ê 0 the displacement must be unchanged as given by equation 1731. In addition, equation 1743 shows that<br />

the magnitude || 2 of the displacement four vector is <strong>in</strong>variant to a Lorentz transformation.<br />

The most general Lorentz transformation between <strong>in</strong>ertial coord<strong>in</strong>ate systems and 0 , <strong>in</strong> relative motion<br />

with velocity v, assum<strong>in</strong>g that the two sets of axes are aligned, and that their orig<strong>in</strong>s overlap when = 0 =0,<br />

is given by the symmetric matrix where<br />

0 = X <br />

(17.32)<br />

This Lorentz transformation of the four vector X components can be written <strong>in</strong> matrix form as<br />

X 0 = λX (17.33)<br />

Assum<strong>in</strong>g that the two sets of axes are aligned, then the elements of the Lorentz transformation are<br />

given by<br />

⎛<br />

⎞<br />

⎛ ⎞<br />

0 − 1 − 2 − 3<br />

⎛ ⎞<br />

X 0 = ⎜ 01<br />

⎟<br />

⎝ 02 ⎠ = − 1 1+( − 1) 2 1<br />

( − 1) 2 1 2<br />

( − 1) 2 1 3<br />

<br />

2<br />

⎜<br />

⎝ − 2 ( − 1) 1 2<br />

1+( − 1) 2<br />

03<br />

2 2<br />

( − 1) 2 2 3 ⎟<br />

2 ⎠ ·<br />

⎜ 1<br />

⎟<br />

⎝ 2 ⎠ (17.34)<br />

− 3 ( − 1) 1 3<br />

( − 1) 2 2 3<br />

1+( − 1) 2 2 3<br />

3<br />

2<br />

where = and = √1− 1 and assum<strong>in</strong>g that the orig<strong>in</strong> of transforms to the orig<strong>in</strong> of 0 at (0 0 0 0).<br />

2<br />

For the case illustrated <strong>in</strong> figure 171 where the correspond<strong>in</strong>g axes of the two frames are parallel and <strong>in</strong><br />

relative motion with velocity <strong>in</strong> the 1 direction, then the Lorentz transformation matrix 1734 reduces to<br />

⎛<br />

⎜<br />

⎝<br />

0<br />

01<br />

02<br />

03<br />

⎞<br />

⎛<br />

⎟<br />

⎠ = ⎜<br />

⎝<br />

− 0 0<br />

− 0 0<br />

0 0 1 0<br />

0 0 0 1<br />

⎞<br />

⎟<br />

⎠ ·<br />

⎛<br />

⎜<br />

⎝<br />

<br />

1<br />

2<br />

3<br />

⎞<br />

⎟<br />

⎠ (17.35)<br />

This Lorentz transformation matrix is called a standard boost s<strong>in</strong>ce it only boosts from one frame to another<br />

parallel frame. In general a rotation matrix also is <strong>in</strong>corporated <strong>in</strong>to the transformation matrix for the<br />

spatial variables.


468 CHAPTER 17. RELATIVISTIC MECHANICS<br />

17.5.2 Four-vector scalar products<br />

Scalar products of vectors and tensors usually are <strong>in</strong>variant to rotations <strong>in</strong> three-dimensional space provid<strong>in</strong>g<br />

an easy way to solve problems. The scalar, or <strong>in</strong>ner, product of two four vectors is def<strong>in</strong>ed by<br />

⎛<br />

⎞ ⎛ ⎞<br />

X · Y = = ¡ 0 1 2 ¢ 1 0 0 0 0<br />

3 · ⎜ 0 −1 0 0<br />

⎟<br />

⎝ 0 0 −1 0 ⎠ ·<br />

⎜ 1<br />

⎟<br />

⎝ 2 ⎠ (17.36)<br />

0 0 0 −1 3<br />

= 0 0 − 1 1 − 2 2 − 3 3<br />

The correct sign of the <strong>in</strong>ner product is obta<strong>in</strong>ed by <strong>in</strong>clusion of the M<strong>in</strong>kowski metric def<strong>in</strong>ed by<br />

that is, it can be represented by the matrix<br />

⎛<br />

≡<br />

⎜<br />

⎝<br />

≡ ê · ê (17.37)<br />

1 0 0 0<br />

0 −1 0 0<br />

0 0 −1 0<br />

0 0 0 −1<br />

⎞<br />

⎟<br />

⎠ (17.38)<br />

The sign convention used <strong>in</strong> the M<strong>in</strong>kowski metric, equation 1738 has been chosen with the time coord<strong>in</strong>ate<br />

() 2 positive which makes () 2 0 for objects mov<strong>in</strong>g at less than the speed of light and corresponds to<br />

be<strong>in</strong>g real. 2<br />

The presence of the M<strong>in</strong>kowski metric matrix, <strong>in</strong> the <strong>in</strong>ner product of four vectors, complicates General<br />

Relativity and thus the E<strong>in</strong>ste<strong>in</strong> convention has been adopted where the components of the contravariant<br />

four-vector X are written with superscripts See also appendix . The correspond<strong>in</strong>g covariant fourvector<br />

components are written with the subscript which is related to the contravariant four-vector<br />

components us<strong>in</strong>g the component of the covariant M<strong>in</strong>kowski metric matrix g That is<br />

3X<br />

= (17.39)<br />

=0<br />

The contravariant metric component is def<strong>in</strong>ed as the component of the <strong>in</strong>verse metric matrix g −1<br />

where<br />

gg −1 = I = g −1 g (17.40)<br />

where I is the four-vector identity matrix. The contravariant components of the four vector can be expressed<br />

<strong>in</strong> terms of the covariant components as<br />

3X<br />

= (17.41)<br />

=0<br />

Thus equations 1739 and 1741 can be used to transform between covariant and contravariant four vectors,<br />

that is, to raise or lower the <strong>in</strong>dex .<br />

The scalar <strong>in</strong>ner product of two four vectors can be written compactly as the scalar product of a covariant<br />

four vector and a contravariant four vector. The M<strong>in</strong>kowski metric matrix can be absorbed <strong>in</strong>to either X or<br />

Y thus<br />

3X 3X<br />

3X<br />

3X<br />

X · Y = = = (17.42)<br />

=0 =0<br />

If this covariant expression is Lorentz <strong>in</strong>variant <strong>in</strong> one coord<strong>in</strong>ate system, then it is Lorentz <strong>in</strong>variant <strong>in</strong> all<br />

coord<strong>in</strong>ate systems obta<strong>in</strong>ed by proper Lorentz transformations.<br />

2 Older textbooks, such as all editions of Marion, and the first two editions of Goldste<strong>in</strong>, use the Euclidean Po<strong>in</strong>caré 4-<br />

dimensional space-time with the imag<strong>in</strong>ary time axis . About half the scientific community, and modern physics textbooks<br />

<strong>in</strong>clud<strong>in</strong>g this textbook and the 3 edition of Goldste<strong>in</strong>, use the Bjorken - Drell + − − −, sign convention given <strong>in</strong> equation<br />

1738 where 0 ≡ and 1 2 3 are the spatial coord<strong>in</strong>ates. The other half of the community, <strong>in</strong>clud<strong>in</strong>g mathematicians<br />

and gravitation physicists, use the opposite − + + + sign convention. Further confusion is caused by a few books that assign<br />

the time axis to be 4 rather than 0 <br />

=0<br />

=0


17.5. GEOMETRY OF SPACE-TIME 469<br />

The scalar <strong>in</strong>ner product of the <strong>in</strong>variant space-time <strong>in</strong>terval is an especially important example.<br />

() 2 ≡ X·X= 2 () 2 − (r) 2 =() 2 −<br />

3X<br />

2 =() 2 (17.43)<br />

This is <strong>in</strong>variant to a Lorentz transformation as can be shown by apply<strong>in</strong>g the Lorentz standard boost<br />

transformation given above. In particular, if 0 is the rest frame of the clock, then the <strong>in</strong>variant space-time<br />

<strong>in</strong>terval is simply given by the proper time <strong>in</strong>terval .<br />

=1<br />

17.5.3 M<strong>in</strong>kowski space-time<br />

Figure 176 illustrates a three-dimensional ¡ 1 2¢ representation of the 4−dimensional space-time diagram<br />

where it is assumed that 3 =0. The fact that the velocity of light has a fixed velocity leads to the<br />

concept of the light cone def<strong>in</strong>ed by the locus of || = .<br />

Inside the light cone<br />

The vertex of the cones represent the present. Locations <strong>in</strong>side<br />

the upper cone represent the future while the past is<br />

represented by locations <strong>in</strong>side the lower cone. Note that<br />

() 2 = 2 () 2 − (r) 2 0 <strong>in</strong>side both the future and past<br />

light cones. Thus the space-time <strong>in</strong>terval ∆ is real and positive<br />

for the future, whereas it is real and negative for the<br />

past relative to the vertex of the light cone. A world l<strong>in</strong>e<br />

is the trajectory a particle follows is a function of time <strong>in</strong><br />

M<strong>in</strong>kowski space. In the <strong>in</strong>terior of the future light cone<br />

∆ 0 and, s<strong>in</strong>ce it is real, it can be asserted unambiguously<br />

that any po<strong>in</strong>t <strong>in</strong>side this forward cone must occur later than<br />

at the vertex of the cone, that is, it is the absolute future.<br />

A Lorentz transformation can rotate M<strong>in</strong>kowski space such<br />

that the axis 0 goes through any po<strong>in</strong>t with<strong>in</strong> this light cone<br />

and then the “world l<strong>in</strong>e” is pure time like. Similarly, any<br />

po<strong>in</strong>t <strong>in</strong>side the backward light cone unambiguously occurred<br />

before the vertex, i.e. it is absolute past.<br />

Outside the light cone<br />

Outside of the light cone, has () 2 = 2 () 2 − (r) 2 0<br />

and thus ∆ is imag<strong>in</strong>ary and is called space like. A spacelike<br />

plane hypersurface <strong>in</strong> spatial coord<strong>in</strong>ates is shown for the<br />

present time <strong>in</strong> the unprimed frame. A rotation <strong>in</strong> M<strong>in</strong>kowski<br />

space can be made to 0 suchthatthespace-likehypersurface<br />

now is tilted relative to the hypersurface shown and thus any<br />

po<strong>in</strong>t outside the light cone can be made to occur later,<br />

simultaneous, or earlier than at the vertex depend<strong>in</strong>g on the<br />

Figure 17.6: The light cone <strong>in</strong> the<br />

1 2 space is def<strong>in</strong>ed by the condition<br />

X · X = 2 2 − 2 =0and divides space-time<br />

<strong>in</strong>to the forward and backward light cones,<br />

with 0 and 0 respectively; the <strong>in</strong>teriors<br />

of the forward and backward light cones<br />

are called absolute future and absolute past.<br />

orientation of the space-like hypersurface. This startl<strong>in</strong>g situation implies that the time order<strong>in</strong>g of two<br />

po<strong>in</strong>ts, each outside the others light cone, can be reversed which has profound implications related to the<br />

concept of simultaneity and the notion of causality.<br />

For the special case of two events ly<strong>in</strong>g on the light cone P 4<br />

2 = 2 2 − ¡ 2 1 + 2 2 + 2 3¢<br />

=0and thus<br />

these events are separated by a light ray travell<strong>in</strong>g at velocity Only events separated by time-like <strong>in</strong>tervals<br />

can be connected causally. The world l<strong>in</strong>e of a particle must lie with<strong>in</strong> its light cone. The division of <strong>in</strong>tervals<br />

<strong>in</strong>to space-like and time-like, because of their <strong>in</strong>variance, is an absolute concept. That is, it is <strong>in</strong>dependent<br />

of the frame of reference.<br />

The concept of proper time can be expanded by consider<strong>in</strong>g a clock at rest <strong>in</strong> frame 0 which is mov<strong>in</strong>g<br />

with uniform velocity with respect to a rest frame . The clock at rest <strong>in</strong> the 0 frame measures the proper


470 CHAPTER 17. RELATIVISTIC MECHANICS<br />

time , then the time observed <strong>in</strong> the fixedframecanbeobta<strong>in</strong>edbylook<strong>in</strong>gatthe<strong>in</strong>terval Because of<br />

the<strong>in</strong>varianceofthe<strong>in</strong>terval, 2 then<br />

2 = 2 2 = 2 2 − £ 2 1 + 2 2 + ¤<br />

2 3<br />

(17.44)<br />

That is,<br />

"<br />

= 1 −<br />

¡<br />

<br />

2<br />

1 + 2 2 + # 1<br />

3¢<br />

2 2<br />

¸ 1<br />

2 2 = <br />

∙1 − 2 2<br />

=<br />

2 <br />

that is = which satisfies the normal expression for time dilation, 178.<br />

17.5.4 Momentum-energy four vector<br />

(17.45)<br />

The previous four-vector discussion can be elegantly exploited us<strong>in</strong>g the covariant M<strong>in</strong>kowski space-time<br />

representation. Separat<strong>in</strong>g the spatial and time of the differential four vector gives<br />

X =( x) (17.46)<br />

Remember that the square of the four-dimensional space-time element of length () 2 is <strong>in</strong>variant (1743),<br />

and is simply related to the proper time element . Thus the scalar product<br />

X·X = 2 = 2 2 = 2 2 − £ 2 1 + 2 2 + 2 ¤<br />

3<br />

(17.47)<br />

Thusthepropertimeisan<strong>in</strong>variant.<br />

The ratio of the four-vector element X and the <strong>in</strong>variant proper time <strong>in</strong>terval is a four-vector called<br />

the four-vector velocity U where<br />

U = X µ<br />

= <br />

x µ<br />

= <br />

x <br />

= <br />

( u) (17.48)<br />

1<br />

where u istheparticlevelocity,and = <br />

(1− 2 ). <br />

The four-vector momentum P can be obta<strong>in</strong>ed 2 from the four-vector velocity by multiply<strong>in</strong>g it by the<br />

scalar rest mass <br />

P = U =( u) (17.49)<br />

However,<br />

= (17.50)<br />

<br />

thus the momentum four vector can be written as<br />

µ <br />

P =<br />

p (17.51)<br />

where the vector p represents the three spatial components of the relativistic momentum. It is <strong>in</strong>terest<strong>in</strong>g to<br />

realize that the Theory of Relativity couples not only the spatial and time coord<strong>in</strong>ates, but also, it couples<br />

their conjugate variables l<strong>in</strong>ear momentum p and total energy, .<br />

An additional feature of this momentum-energy four vector P, is that the scalar <strong>in</strong>ner product P · P is<br />

<strong>in</strong>variant to Lorentz transformations and equals () 2 <strong>in</strong> the rest frame<br />

P · P =<br />

3X<br />

=0 =0<br />

3X<br />

=<br />

3X<br />

=0 =0<br />

3X<br />

=( )2 − |p| 2 = 2 2 (17.52)<br />

which leads to the well-known equation<br />

2 = 2 2 + 0 2 (17.53)<br />

The Lorentz transformation matrix canbeappliedtoP<br />

P 0 = λP (17.54)<br />

The Lorentz <strong>in</strong>variant four-vector representation ³ is illustrated by apply<strong>in</strong>g the Lorentz transformation<br />

shown <strong>in</strong> figure 171, whichgives, 0 1 = 1 − ¡ ¢<br />

2<br />

´<br />

, 0 2 = 2 , 0 3 = 3 ,and 0 = ( − 1 ).


17.6. LORENTZ-INVARIANT FORMULATION OF LAGRANGIAN MECHANICS 471<br />

17.6 Lorentz-<strong>in</strong>variant formulation of Lagrangian mechanics<br />

17.6.1 Parametric formulation<br />

The Lagrangian and Hamiltonian formalisms <strong>in</strong> classical mechanics are based on the Newtonian concept<br />

of absolute time which serves as the system evolution parameter <strong>in</strong> Hamilton’s Pr<strong>in</strong>ciple. This approach<br />

violates the Special Theory of Relativity. The extended Lagrangian and Hamiltonian formalism is a parametric<br />

approach, pioneered by Lanczos[La49], that <strong>in</strong>troduces a system evolution parameter that serves<br />

as the <strong>in</strong>dependent variable <strong>in</strong> the action <strong>in</strong>tegral, and all the space-time variables ()() are dependent<br />

on the evolution parameter . This extended Lagrangian and Hamiltonian formalism renders it to a form<br />

that is compatible with the Special Theory of Relativity. The importance of the Lorentz-<strong>in</strong>variant extended<br />

formulation of Lagrangian and Hamiltonian mechanics has been recognized for decades.[La49, Go50, Sy60]<br />

Recently there has been a resurgence of <strong>in</strong>terest <strong>in</strong> the extended Lagrangian and Hamiltonian formalism<br />

stimulated by the papers of Struckmeier[Str05, Str08] and this formalism has featured prom<strong>in</strong>ently <strong>in</strong> recent<br />

textbooks by Johns[Jo05] and Gre<strong>in</strong>er[Gr10]. This parametric approach develops manifestly-covariant Lagrangian<br />

and Hamiltonian formalisms that treat equally all 2+1 space-time canonical variables. It provides<br />

a plausible manifestly-covariant Lagrangian for the one-body system, but serious problems exist extend<strong>in</strong>g<br />

this to the -body system when 1. Generaliz<strong>in</strong>g the Lagrangian and Hamiltonian formalisms <strong>in</strong>to the<br />

doma<strong>in</strong> of the Special Theory of Relativity is of fundamental importance to physics, while the parametric<br />

approach gives <strong>in</strong>sight <strong>in</strong>to the philosophy underly<strong>in</strong>g use of variational methods <strong>in</strong> classical mechanics. 3<br />

In conventional Lagrangian mechanics, the equations of motion for the generalized coord<strong>in</strong>ates are<br />

derived by m<strong>in</strong>imiz<strong>in</strong>g the action <strong>in</strong>tegral, that is, Hamilton’s Pr<strong>in</strong>ciple.<br />

(q ˙q) =<br />

Z <br />

<br />

(q() ˙q()) =0 (17.55)<br />

where (q() ˙q()) denotes the conventional Lagrangian. This approach implicitly assumes the Newtonian<br />

concept of absolute time which is chosen to be the <strong>in</strong>dependent variable that characterizes the evolution<br />

parameter of the system. The actual path [q() ˙q()] the system follows is def<strong>in</strong>ed by the extremum of the<br />

action <strong>in</strong>tegral (q ˙q) which leads to the correspond<strong>in</strong>g Euler-Lagrange equations. This assumption is<br />

contrary to the Theory of Relativity which requires that the space and time variables be treated equally,<br />

that is, the Lagrangian formalism must be covariant.<br />

17.6.2 Extended Lagrangian<br />

Lanczos[La49] proposed mak<strong>in</strong>g the Lagrangian covariant by <strong>in</strong>troduc<strong>in</strong>g a general evolution parameter <br />

and treat<strong>in</strong>g the time as a dependent variable () on an equal foot<strong>in</strong>g with the configuration space variables<br />

() That is, the time becomes a dependent variable 0 () =() similar to the spatial variables ()<br />

where 1 ≤ ≤ . The dynamical system then is described as motion conf<strong>in</strong>ed to a hypersurface with<strong>in</strong> an<br />

extended space where the value of the extended Hamiltonian and the evolution parameter constitute an<br />

additional pair of canonically conjugate variables <strong>in</strong> the extended space. That is, the canonical momentum<br />

0 correspond<strong>in</strong>g to 0 = is 0 = <br />

similar to the momentum-energy four vector, equation 1751.<br />

An extended Lagrangian L(q() q()<br />

<br />

() ()<br />

<br />

) can be def<strong>in</strong>ed which can be written compactly as<br />

L( () ()<br />

<br />

) where the <strong>in</strong>dex 0 ≤ ≤ denotes the entire range of space-time variables.<br />

This extended Lagrangian can be used <strong>in</strong> an extended action functional S(q q <br />

<br />

<br />

<br />

) to give an extended<br />

version of Hamilton’s Pr<strong>in</strong>ciple 4 S(q q <br />

<br />

<br />

)= Z <br />

<br />

L( () ()<br />

) =0 (17.56)<br />

<br />

3 Chapters 176 and 177 reproduce the Struckmeier presentation.[Str08]<br />

4 These formula <strong>in</strong>volve total and partial derivatives with respect to both time, and parameter . For clarity, the derivatives<br />

are written out <strong>in</strong> full because Lanczos[La49] and Johns[Jo05] use the opposite convention for the dot and prime superscripts<br />

as abbreviations for the differentials with respect to and . The blackboard bold format is used to designate the extended<br />

versions of the action , Lagrangian and Hamiltonian .


472 CHAPTER 17. RELATIVISTIC MECHANICS<br />

The conventional action andextendedactionS, address alternate characterizations of the same underly<strong>in</strong>g<br />

physical system, and thus the action pr<strong>in</strong>ciple implies that = S =0must hold simultaneously. That is,<br />

<br />

Z <br />

<br />

(q q <br />

)<br />

<br />

= Z <br />

<br />

L(q q <br />

) (17.57)<br />

<br />

As discussed <strong>in</strong> chapter 93 there is a cont<strong>in</strong>uous spectrum of equivalent gauge-<strong>in</strong>variant Lagrangians for<br />

which the Euler-Lagrange equations lead to identical equations of motion. Equation 1757 is satisfied if the<br />

conventional and extended Lagrangians are related by<br />

L(q q <br />

<br />

)=(qq <br />

)<br />

+ Λ(q)<br />

(17.58)<br />

<br />

where Λ(q) is a cont<strong>in</strong>uous function of q and that has cont<strong>in</strong>uous second derivatives. It is acceptable to<br />

assume that Λ(q)<br />

<br />

=0, then the extended and conventional Lagrangians have a unique relation requir<strong>in</strong>g<br />

no simultaneous transformation of the dynamical variables. That is, assume<br />

L(q q <br />

<br />

)=(qq <br />

) (17.59)<br />

<br />

Note that the time derivative of q can be expressed <strong>in</strong> terms of the derivatives by<br />

q<br />

= q<br />

(17.60)<br />

<br />

Thus, for a conventional Lagrangian with variables, the correspond<strong>in</strong>g extended Lagrangian is a function<br />

of +1 variables while the conventional and extended Lagrangians are related us<strong>in</strong>g equations 1759 and<br />

1760.<br />

The derivatives of the relation between the extended and conventional Lagrangians lead to<br />

L<br />

= <br />

(17.61)<br />

<br />

L<br />

= <br />

(17.62)<br />

<br />

L<br />

<br />

³ ´ = ³ ´ (17.63)<br />

<br />

<br />

<br />

<br />

L<br />

X<br />

¡ <br />

¢ = − ´ <br />

(17.64)<br />

<br />

<br />

<br />

where 1 ≤ ≤ s<strong>in</strong>ce the =0time derivatives are written explicitly <strong>in</strong> equations 1762 1764.<br />

Equations 1763 — 1764, summed over the extended range 0 ≤ ≤ of time and spatial dynamical<br />

variables, imply<br />

X<br />

=0<br />

<br />

L<br />

³<br />

<br />

<br />

µ <br />

<br />

´<br />

<br />

= <br />

− X<br />

=1<br />

<br />

=1<br />

<br />

³<br />

<br />

<br />

³<br />

<br />

<br />

´ <br />

<br />

<br />

<br />

+ X<br />

=1<br />

<br />

<br />

³<br />

<br />

<br />

´ <br />

= L (17.65)<br />

Equation 1765 can be written <strong>in</strong> the form<br />

X<br />

½<br />

L<br />

L− ´ 6≡=<br />

= 0 if L is not homogeneous <strong>in</strong> <br />

<br />

≡ 0 if L is homogeneous <strong>in</strong> <br />

<br />

=0<br />

³<br />

<br />

<br />

(17.66)<br />

If the extended Lagrangian L(q q <br />

<br />

<br />

<br />

<br />

) is homogeneous to first order <strong>in</strong> the +1 variables<br />

<br />

, then Euler’s<br />

theorem on homogeneous functions trivially implies the relation given <strong>in</strong> equation 1766. Struckmeier[Str08]<br />

identified a subtle but important po<strong>in</strong>t that if L is not homogeneous <strong>in</strong> <br />

<br />

then equation 1766 is not an<br />

identity but is an implicit equation that is always satisfied as the system evolves accord<strong>in</strong>g to the solution<br />

of the extended Euler-Lagrange equations. Then equation 1759 is satisfied without it be<strong>in</strong>g a homogeneous<br />

form <strong>in</strong> the +1 velocities <br />

<br />

. This <strong>in</strong>troduces a new class of non-homogeneous Lagrangians. The relativistic<br />

free particle, discussed <strong>in</strong> example 175 is a case of a non-homogeneous extended Lagrangian.


17.6. LORENTZ-INVARIANT FORMULATION OF LAGRANGIAN MECHANICS 473<br />

17.6.3 Extended generalized momenta<br />

The generalized momentum is def<strong>in</strong>ed by<br />

=<br />

´ (17.67)<br />

<br />

Assume that the def<strong>in</strong>itions of the extended Lagrangian L, and the extended Hamiltonian H, are related<br />

by a Legendre transformation, and are based on variational pr<strong>in</strong>ciples, analogous to the relation that exists<br />

between the conventional Lagrangian and Hamiltonian . The Legendre transformation requires def<strong>in</strong><strong>in</strong>g<br />

the extended generalized (canonical) momentum-energy four vector P()= ( ()<br />

<br />

p()). The momentum<br />

components of the momentum-energy four vector P()= ( ()<br />

<br />

p()) are given by the 1 ≤ ≤ components<br />

us<strong>in</strong>g equation 1763<br />

³<br />

<br />

<br />

³<br />

<br />

<br />

() =<br />

L ´ = ´ (17.68)<br />

<br />

The =0component of the momentum-energy four vector can be derived by recogniz<strong>in</strong>g that the right-hand<br />

side of equation 1764 is equal to −( ). That is, the correspond<strong>in</strong>g generalized momentum 0 that<br />

is conjugate to 0 = is given by<br />

⎛<br />

⎞<br />

Ã<br />

0 =<br />

L ´ = 1<br />

<br />

³<br />

0<br />

<br />

!<br />

L<br />

¡ ¢ = 1 ⎝ −<br />

<br />

<br />

<br />

X<br />

=1<br />

<br />

³<br />

<br />

<br />

17.6.4 Extended Lagrange equations of motion<br />

<br />

³<br />

<br />

<br />

´ <br />

<br />

⎠ = − ( )<br />

<br />

(17.69)<br />

By direct analogy with the non-relativistic action <strong>in</strong>tegral 1755 the extremum for the relativistic action<br />

<strong>in</strong>tegral S(q q <br />

<br />

<br />

<br />

) is obta<strong>in</strong>ed us<strong>in</strong>g the Euler-Lagrange equations derived from equation 1756 where the<br />

<strong>in</strong>dependent variable is . This implies that for 0 ≤ ≤ <br />

⎛ ⎞<br />

<br />

⎝<br />

L ´ ⎠ − L<br />

X<br />

<br />

= <br />

Q =<br />

<br />

<br />

+ <br />

<br />

(17.70)<br />

<br />

³<br />

<br />

<br />

where the extended generalized force Q <br />

shown on the right-hand side of equation 1770 accounts for all<br />

forces not <strong>in</strong>cluded <strong>in</strong> the potential energy term <strong>in</strong> the Lagrangian. The extended generalized force Q <br />

can<br />

be factored <strong>in</strong>to two terms as discussed <strong>in</strong> chapter 6, equation660. The Lagrange multiplier term <strong>in</strong>cludes<br />

1 ≤ ≤ holonomic constra<strong>in</strong>t forces where the holonomic constra<strong>in</strong>ts, which do no work, are expressed<br />

<strong>in</strong> terms of the algebraic equations of holonomic constra<strong>in</strong>t . The <br />

term <strong>in</strong>cludes the rema<strong>in</strong><strong>in</strong>g<br />

constra<strong>in</strong>t forces and generalized forces that are not <strong>in</strong>cluded <strong>in</strong> the Lagrange multiplier term or the potential<br />

energy term of the Lagrangian.<br />

For the case where =0,s<strong>in</strong>ce 0 = , thenequation1770 reduces to<br />

<br />

<br />

à !<br />

L<br />

¡ ¢<br />

<br />

<br />

− L<br />

= X<br />

=1<br />

=1<br />

<br />

<br />

<br />

− X<br />

=1<br />

<br />

<br />

<br />

(17.71)<br />

These Euler-Lagrange equations of motion 1770 1771 determ<strong>in</strong>e the 1 ≤ ≤ generalized coord<strong>in</strong>ates<br />

() plus 0 = () <strong>in</strong> terms of the <strong>in</strong>dependent variable .<br />

If the holonomic equations of constra<strong>in</strong>t are time <strong>in</strong>dependent, that is <br />

<br />

=0and if Q <br />

0 =0,then<br />

the =0term of the Euler-Lagrange equations simplifies to<br />

à !<br />

L<br />

¡ ¢ − L =0 (17.72)<br />

<br />

<br />

<br />

One <strong>in</strong>terpretation is to select to be primary. Then L is derived from us<strong>in</strong>g equation 1759 and L<br />

must satisfy the identity given by equation 1766 while the Euler-Lagrange equations conta<strong>in</strong><strong>in</strong>g <br />

<br />

yield an<br />

identity which implies that does not provide an equation of motion <strong>in</strong> terms of (). Conversely, if L is


474 CHAPTER 17. RELATIVISTIC MECHANICS<br />

chosen to be primary, then L is no longer a homogeneous function and equation 1766 serves as a constra<strong>in</strong>t<br />

on the motion that can be used to deduce , while <br />

<br />

yields a non-trivial equation of motion <strong>in</strong> terms of<br />

(). In both cases the occurrence of a constra<strong>in</strong>t surface results from the fact that the extended space has<br />

2 +2 variables to describe 2 +1 degrees of freedom, that is, one more degree of freedom than required for<br />

the actual system.<br />

17.5 Example: Lagrangian for a relativistic free particle<br />

The standard Lagrangian = − is not Lorentz <strong>in</strong>variant. The extended Lagrangian ( q <br />

<br />

<br />

)<br />

<strong>in</strong>troduces the <strong>in</strong>dependent variable which treats both the space variables () and time variable 0 = ()<br />

equally. This can be achieved by def<strong>in</strong><strong>in</strong>g the non-standard Lagrangian<br />

"<br />

L(q q<br />

µ #<br />

2<br />

<br />

<br />

)=1 1 q<br />

2 2 2 − ( <br />

)2 − 1<br />

()<br />

The constant third term <strong>in</strong> the bracket is <strong>in</strong>cluded to ensure that the extended Lagrangian converges to the<br />

standard Lagrangian <strong>in</strong> the limit <br />

→ 1.<br />

Note that the extended Lagrangian () is not homogeneous to first order <strong>in</strong> the velocities q<br />

<br />

as is required.<br />

Equation 1766 must be used to ensure that equation () is homogeneous. That is, it must satisfy the<br />

constra<strong>in</strong>t relation<br />

µ 2 <br />

− 1 µ 2 q<br />

2 − 1=0 ()<br />

<br />

Insert<strong>in</strong>g () <strong>in</strong>to the extended Lagrangian () yields that the square bracket <strong>in</strong> equation must equal 2.<br />

Thus<br />

|L| = 1 2 2 [−2] = − 2<br />

()<br />

The constra<strong>in</strong>t equation () implies that<br />

s<br />

<br />

= 1 − 1 µ 2 q<br />

2 = 1 <br />

()<br />

Us<strong>in</strong>g equation () gives that the relativistic Lagrangian is<br />

= L = −2 <br />

= − 2 q1 − 2 ()<br />

Equation () is the conventional relativistic Lagrangian derived by assum<strong>in</strong>g that the system evolution parameter<br />

is transformed to be along the world l<strong>in</strong>e where the <strong>in</strong>variant length replaces the proper time<br />

<strong>in</strong>terval<br />

= = <br />

()<br />

<br />

The def<strong>in</strong>ition of the generalized (canonical) momentum<br />

= <br />

˙ <br />

= ˙ <br />

()<br />

leads to the relativistic expression for momentum given <strong>in</strong> equation 1721.<br />

The relativistic Lagrangian is an important example of a non-standard Lagrangian. Equation () does not<br />

equal the difference between the k<strong>in</strong>etic and potential energies, that is, the relativistic expression for k<strong>in</strong>etic<br />

energy is given by 1728 to be<br />

=( − 1) 2<br />

()<br />

The non-standard relativistic Lagrangian () can be used with the Euler-Lagrange equations to derive the<br />

second-order equations of motion for both relativistic and non-relativistic problems with<strong>in</strong> the Special Theory<br />

of Relativity.


17.6. LORENTZ-INVARIANT FORMULATION OF LAGRANGIAN MECHANICS 475<br />

17.6 Example: Relativistic particle <strong>in</strong> an external electromagnetic field<br />

A charged particle mov<strong>in</strong>g at relativistic speed <strong>in</strong> an external electromagnetic field provides an example<br />

of the use of the relativistic Lagrangian.<br />

In the discussion of classical mechanics it was shown that the velocity-dependent Lorentz force can be<br />

absorbed <strong>in</strong>to the scalar electric potential Φ plus the vector magnetic potential A. That is, the potential<br />

energy is given by equation 76 to be = (Φ − A · v) Includ<strong>in</strong>g this <strong>in</strong> the Lagrangian, 1771 gives<br />

= − 2<br />

<br />

− = −2 q<br />

1 − 2 − Φ + A · v<br />

The three spatial partial derivatives can be written <strong>in</strong> vector notation as<br />

<br />

r = −∇Φ + ∇(v · A)<br />

()<br />

and the generalized momentum is given by<br />

p = = v + A<br />

v<br />

which is identical to the non-relativistic answer given by equation 76. That is, it <strong>in</strong>cludes the momentum of<br />

the electromagnetic field plus the classical l<strong>in</strong>ear momentum of the mov<strong>in</strong>g particle.<br />

The total time derivative of the generalized momentum is<br />

p<br />

= <br />

where the last term is given by the cha<strong>in</strong> rule<br />

µ <br />

v<br />

<br />

= A<br />

(v)+<br />

<br />

A<br />

= A +(v · ∇)A<br />

()<br />

Us<strong>in</strong>g equations <strong>in</strong> the Euler-Lagrange equation gives<br />

<br />

<br />

µ <br />

v<br />

<br />

= <br />

r<br />

A<br />

(v)+ = −∇Φ + ∇(v · A)<br />

<br />

Collect<strong>in</strong>g terms and us<strong>in</strong>g the well-known vector-product identity, plus the def<strong>in</strong>ition B = ∇ × A gives<br />

∙<br />

<br />

(v) = − ∇Φ − A ¸<br />

+ [∇(v · A) − (v · ∇)A]<br />

<br />

= −<br />

∙<br />

∇Φ − A<br />

<br />

F = [E + v × B]<br />

¸<br />

+ [v × ∇ × A]<br />

If we adopt the def<strong>in</strong>ition that the relativistic canonical momentum is = then the left hand side is<br />

the relativistic force while the right-hand side is the well-known Lorentz force of electromagnetism. Thus<br />

the extended Lagrangian formulation correctly reproduces the well-known Lorentz force for a charged particle<br />

mov<strong>in</strong>g <strong>in</strong> an electromagnetic field.<br />

()


476 CHAPTER 17. RELATIVISTIC MECHANICS<br />

17.7 Lorentz-<strong>in</strong>variant formulations of Hamiltonian mechanics<br />

17.7.1 Extended canonical formalism<br />

A Lorentz-<strong>in</strong>variant formulation of Hamiltonian mechanics can be developed that is built upon the extended<br />

Lagrangian formalism assum<strong>in</strong>g that the Hamiltonian and Lagrangian are related by a Legendre transformation.<br />

That is,<br />

(q p)=<br />

where the generalized momentum is def<strong>in</strong>ed by<br />

X<br />

=1<br />

<br />

<br />

<br />

³<br />

<br />

<br />

q<br />

− (q ) (17.73)<br />

<br />

=<br />

´ (17.74)<br />

<br />

Struckmeier[Str08] assumes that the def<strong>in</strong>itions of the extended Lagrangian L, and the extended Hamiltonian<br />

H, are related by a Legendre transformation, and are based on variational pr<strong>in</strong>ciples, analogous to the<br />

relation that exists between the conventional Lagrangian and Hamiltonian . The Legendre transformation<br />

requires def<strong>in</strong><strong>in</strong>g the extended generalized (canonical) momentum-energy four vector P()= ( ()<br />

<br />

p()).<br />

The momentum components of the momentum-energy four vector P()= ( ()<br />

<br />

p()) are given by the 1 ≤<br />

≤ components us<strong>in</strong>g either the conventional or the extended Lagrangians as given <strong>in</strong> equation 1768<br />

³<br />

<br />

<br />

³<br />

<br />

<br />

() =<br />

L ´ = ´ (1768)<br />

<br />

The =0component of the momentum-energy four vector is given by equation 1769<br />

à !<br />

0 = 1 L<br />

¡ ¢ = − ( )<br />

= − E()<br />

(17.75)<br />

<br />

<br />

<br />

<br />

where E() represents the <strong>in</strong>stantaneous generalized energy of the conventional Hamiltonian at the po<strong>in</strong>t <br />

but not the functional form of (q() p()()). Thatis<br />

E() 6≡ =(q() p()()) (17.76)<br />

Note that E() does not give the function (q p). Equations1768 and 1769 give that<br />

0 () =− E()<br />

(17.77)<br />

<br />

The extended Hamiltonian H(q pE()), <strong>in</strong>anextendedphasespace,canbedef<strong>in</strong>ed by the Legendre<br />

transformation and the four-vector P to be<br />

H(q pE()) = (P·q) − L(q q <br />

<br />

) (17.78)<br />

X<br />

µ <br />

<br />

= − L(q q <br />

<br />

)<br />

=<br />

=0<br />

X<br />

=1<br />

<br />

µ <br />

<br />

<br />

<br />

− E <br />

− L(qq <br />

<br />

<br />

) (17.79)<br />

where the 0 term has been written explicitly as −E <br />

<br />

<strong>in</strong> equation 1779. The extended Hamiltonian<br />

H((q pE()) can carry all the <strong>in</strong>formation on the dynamical system that is carried by the extended<br />

Lagrangian L(q q <br />

<br />

<br />

<br />

) if the Hesse matrix is non-s<strong>in</strong>gular. That is, if<br />

⎛<br />

⎞<br />

det ⎝<br />

2 L<br />

³ ´ ³ ´ ⎠ 6= 0 (17.80)


17.7. LORENTZ-INVARIANT FORMULATIONS OF HAMILTONIAN MECHANICS 477<br />

If the extended Lagrangian L(q q <br />

<br />

<br />

<br />

<br />

) is not homogeneous <strong>in</strong> the +1 velocities<br />

<br />

, then the extended<br />

set of Euler-Lagrange equations 1772 is not redundant. Thus equation 1766 is not an identity but it can be<br />

regarded as an implicit equation that is always satisfied by the extended set of Euler-Lagrange equations. As<br />

a result, the Legendre transformation to an extended Hamiltonian exists. That is, equation 1766 is identical<br />

totheLegendretransformforH((q pE()) which was shown to equal zero. Therefore<br />

H(q() p()() E()) = 0 (17.81)<br />

which means that the extended Hamiltonian H((q pE()) directly def<strong>in</strong>es the restricted hypersurface on<br />

which the particle motion is conf<strong>in</strong>ed.<br />

The extended canonical equations of motion, derived us<strong>in</strong>g the extended Hamiltonian H(q() p()() E())<br />

with the usual Hamiltonian mechanics relations, are:<br />

H<br />

= <br />

(17.82)<br />

<br />

H<br />

= − <br />

(17.83)<br />

<br />

H<br />

= E<br />

(17.84)<br />

<br />

H<br />

= − <br />

(17.85)<br />

E <br />

These canonical equations give that the total derivative of H((q() p()() E()) with respect to is<br />

H<br />

<br />

= H <br />

+ H <br />

+ H <br />

+ H E<br />

E <br />

= <br />

− <br />

+ E <br />

− E<br />

=0 (17.86)<br />

<br />

That is, <strong>in</strong> contrast to the total time derivative of (q p), thetotal derivative of the extended Hamiltonian<br />

H((q() p()() E()) always vanishes, that is, H(q() p()() E()) is autonomous which is ideal<br />

for use with Hamilton’s equations of motion. The constra<strong>in</strong>ts give that H(q() p()() E()) = 0,(equation<br />

1781) and <br />

<br />

=0,(equation1786) imply<strong>in</strong>g that the correlation between the extended and conventional<br />

Hamiltonians is given by<br />

H((q() p()() E()) =<br />

=<br />

=<br />

X<br />

=1<br />

X<br />

=1<br />

X<br />

=1<br />

µ <br />

<br />

− E <br />

<br />

<br />

µ <br />

<br />

<br />

<br />

− E <br />

− L(qq <br />

− (qq <br />

µ "<br />

<br />

<br />

− E <br />

+<br />

<br />

<br />

) (17.87)<br />

)<br />

<br />

<br />

(q p) −<br />

X<br />

=1<br />

µ # <br />

<br />

<br />

<br />

<br />

(17.88)<br />

(17.89)<br />

= ((q p) − E) =0 (17.90)<br />

<br />

s<strong>in</strong>ce only the term with =0does not cancel <strong>in</strong> equation 1779. Equations1781 and 1790 give that both the<br />

left and right-hand sides of equation 1790 are zero while equation 1786 implies that H((q() p()() E())<br />

is a constant of motion, that is, is a cyclic variable for H((q() p()() E()). Formally one can consider<br />

the extended Hamiltonian is a constant which equals zero<br />

H(q pE()) = E() =0 (17.91)<br />

Equations 1784 1785 imply that (E) form a pair of canonically conjugate variables <strong>in</strong> addition to the<br />

newly-<strong>in</strong>troduced canonically-conjugate variables (E()). Equation 1790 shows that the motion <strong>in</strong> the<br />

2 +2 extended phase space is constra<strong>in</strong>ed to the surface reflect<strong>in</strong>g the fact that the observed system has<br />

one less degree of freedom than used by the extended Hamiltonian.<br />

In summary, the Lorentz-<strong>in</strong>variant extended canonical formalism leads to Hamilton’s first-order equations<br />

of motion <strong>in</strong> terms of derivatives with respect to where is related to the proper time for a relativistic<br />

system.


478 CHAPTER 17. RELATIVISTIC MECHANICS<br />

17.7.2 Extended Poisson Bracket representation<br />

Struckmeier[Str08] <strong>in</strong>vestigated the usefulness of the extended formalism when applied to the Poisson bracket<br />

representation of Hamiltonian mechanics. The extended Poisson bracket for two differentiable functions <br />

and is def<strong>in</strong>ed as<br />

X<br />

µ <br />

[[ ]] =<br />

− <br />

<br />

− <br />

+ <br />

(17.92)<br />

<br />

=1<br />

As for the conventional Poisson bracket discussed <strong>in</strong> chapter 15, the extended Poisson also leads to the<br />

fundamental Poisson bracket relations<br />

££<br />

¤¤ ££<br />

=0 [[ ]] = 0<br />

¤¤<br />

= <br />

<br />

(17.93)<br />

where =0 1 . These are identical to the non-extended fundamental Poisson brackets.<br />

The discussion of observables <strong>in</strong> Hamiltonian mechanics <strong>in</strong> chapter 1525 can be trivially expanded to<br />

the extended Poisson bracket representation. In particular, the total derivative of the function is given<br />

by<br />

<br />

= +[[ H]] (17.94)<br />

<br />

If commutes with the extended Hamiltonian, that is, the Poisson bracket equals zero, and if <br />

<br />

<br />

=0. That is, the observable is a constant of motion.<br />

Substitute the fundamental variables for gives<br />

=0,then<br />

<br />

=[[ H]] = − H<br />

<br />

<br />

=[[ H]] = H<br />

<br />

(17.95)<br />

where =0 1 . These are Hamilton’s extended canonical equations of motion expressed <strong>in</strong> terms of<br />

the system evolution parameter . The extended Poisson bracket representation is a trivial extension of the<br />

conventional canonical equations presented <strong>in</strong> chapter 153.<br />

17.7.3 Extended canonical transformation and Hamilton-Jacobi theory<br />

Struckmeier[Str08] presented plausible extended versions of canonical transformation and Hamilton-Jacobi<br />

theories that can be used to provide a Lorentz-<strong>in</strong>variant formulation of Hamiltonian mechanics for relativistic<br />

one-body systems. A detailed description can be found <strong>in</strong> Struckmeier[Str08]. 5<br />

17.7.4 Validity of the extended Hamilton-Lagrange formalism<br />

It has been shown that the extended Lagrangian and Hamiltonian formalism, based on the parametric model<br />

of Lanczos[La49], leads to a plausible manifestly-covariant approach for the one-body system. The general<br />

features developed for handl<strong>in</strong>g Lagrangian and Hamiltonian mechanics carry over to the Special Theory<br />

of Relativity assum<strong>in</strong>g the use of a non-standard, extended Lagrangian or Hamiltonian. This expansion of<br />

the range of validity of the well-known Hamiltonian and Lagrangian mechanics <strong>in</strong>to the relativistic doma<strong>in</strong><br />

is important, and reduces any Lorentz transformation to a canonical transformation. The validity of this<br />

extended Hamilton-Lagrange formalism has been criticized, and problems exist extend<strong>in</strong>g this approach to<br />

the -body system for 1. For example, as discussed by Goldste<strong>in</strong>[Go50] and Johns[Jo05], each of<br />

the mov<strong>in</strong>g bodies have their own world l<strong>in</strong>es and momenta. Def<strong>in</strong><strong>in</strong>g the total momentum P requires<br />

know<strong>in</strong>g simultaneously the momenta of the <strong>in</strong>dividual bodies, but simultaneity is body dependent and<br />

thus even the total momentum is not a simple four vector. A general method is required that will allow<br />

us<strong>in</strong>g a manifestly-covariant Lagrangian or Hamiltonian for the -body system. For the one-body system,<br />

the extended Hamilton-Lagrange formalism provides a powerful and logical approach to exploit analytical<br />

mechanics <strong>in</strong> the relativistic doma<strong>in</strong> that reta<strong>in</strong>s the form of the conventional Lagrangian/Hamiltonian<br />

formalisms. Note that Noether’s theorem relat<strong>in</strong>g energy and time is readily apparent us<strong>in</strong>g the extended<br />

formalism.<br />

5 Note that Gre<strong>in</strong>er[Gr10] <strong>in</strong>cludes a reproduction of the Struckmeier paper[Str08].


17.7. LORENTZ-INVARIANT FORMULATIONS OF HAMILTONIAN MECHANICS 479<br />

17.7 Example: The Bohr-Sommerfeld hydrogen atom<br />

The classical relativistic hydrogen atom was first solved by Sommerfeld <strong>in</strong> 1916. Sommerfeld used Bohr’s<br />

“old quantum theory” plus Hamiltonian mechanics to make an important step <strong>in</strong> the development of quantum<br />

mechanics by obta<strong>in</strong><strong>in</strong>g the first-order expressions for the f<strong>in</strong>e structure of the hydrogen atom. As <strong>in</strong> the<br />

non-relativistic case, the motion is conf<strong>in</strong>ed to a plane allow<strong>in</strong>g use of planar polar coord<strong>in</strong>ates. Thus the<br />

relativistic Lagrangian is given by<br />

s<br />

− = −2 1 − ˙2 + 2 ˙ 2<br />

= − 2<br />

<br />

The canonical momenta are given by<br />

= <br />

˙ = 2 ˙<br />

= = ˙<br />

˙<br />

˙ = <br />

=0<br />

˙ = <br />

= ˙ 2 + 2<br />

2<br />

2<br />

+ 2<br />

<br />

As for the non-relativistic case, is a cyclic variable and thus the<br />

angular momentum = 2 ˙ is conserved.<br />

The relativistic Hamiltonian for the Coulomb potential between an<br />

electron and the proton, assum<strong>in</strong>g that the motion is conf<strong>in</strong>ed to a<br />

plane, which allows use of planar polar coord<strong>in</strong>ates, leads to<br />

The advance of the perihelion of<br />

r<br />

bound orbits due to the dependence<br />

of the relativistic mass on velocity.<br />

= 2 2 + 2 2<br />

2 + 2 4 − 2<br />

<br />

The same equations of motion are obta<strong>in</strong>ed us<strong>in</strong>g Hamiltonian mechanics, that is:<br />

˙ = =<br />

<br />

2<br />

˙ = = <br />

<br />

˙ = − <br />

=0<br />

˙ = − <br />

= ˙ 2 + 2<br />

2<br />

The radial dependence can be solved us<strong>in</strong>g either Lagrangian or Hamiltonian mechanics, but the solution<br />

is non-trivial. Us<strong>in</strong>g the same techniques applied to solve Kepler’s problem, leads to the radial solution<br />

=<br />

<br />

1+ cos[Γ( − 0 ]<br />

Γ =<br />

s<br />

1 − 4<br />

2 2 <br />

= 2 Γ 2 2 <br />

2 <br />

=<br />

s<br />

1+ Γ2 (1 − 2 4<br />

2 )<br />

1 − Γ 2<br />

The apses are m<strong>in</strong> =<br />

<br />

(1+) for Γ( − 0)=0 2 4 and max = <br />

(1−) for Γ( − 0)= 3. The<br />

perihelion advances between cycles due to the change <strong>in</strong> relativistic mass dur<strong>in</strong>g the trajectory as shown <strong>in</strong><br />

the adjacent figure.Thisprecessionleadstothef<strong>in</strong>e structure observed <strong>in</strong> the optical spectra of the hydrogen<br />

atom. The same precession of the perihelion occurs for planetary motion, however, there is a comparable<br />

size effect due to gravity that requires use of general relativity to compute the trajectories.


480 CHAPTER 17. RELATIVISTIC MECHANICS<br />

17.8 The General Theory of Relativity<br />

E<strong>in</strong>ste<strong>in</strong>’s General Theory of Relativity expands the scope of relativistic mechanics to <strong>in</strong>clude non-<strong>in</strong>ertial<br />

accelerat<strong>in</strong>g frames plus a unified theory of gravitation. That is, the General Theory of Relativity <strong>in</strong>corporates<br />

both the Special Theory of Relativity as well as Newton’s Law of Universal Gravitation. It provides<br />

a unified theory of gravitation that is a geometric property of space and time. In particular, the curvature<br />

of space-time is directly related to the four-momentum of matter and radiation. Unfortunately, E<strong>in</strong>ste<strong>in</strong>’s<br />

equations of general relativity are nonl<strong>in</strong>ear partial differential equations that are difficult to solve exactly,<br />

and the theory requires knowledge of Riemannian geometry that goes beyond the scope of this book. The<br />

follow<strong>in</strong>g summarizes the fundamental variational concepts underly<strong>in</strong>g the theory, and the experimental<br />

evidence <strong>in</strong> support of the General Theory of Relativity.<br />

17.8.1 The fundamental concepts<br />

E<strong>in</strong>ste<strong>in</strong> <strong>in</strong>corporated the follow<strong>in</strong>g concepts <strong>in</strong> the General Theory of Relativity.<br />

Mach’s pr<strong>in</strong>ciple:<br />

The 1883 work “The Science of <strong>Mechanics</strong>” by the philosopher/physicist, Ernst Mach, criticized Newton’s<br />

concept of an absolute frame of reference, and suggested that local physical laws are determ<strong>in</strong>ed by the largescale<br />

structure of the universe. Mach’s Pr<strong>in</strong>ciple assumes that local motion of a rotat<strong>in</strong>g frame is determ<strong>in</strong>ed<br />

by the large-scale distribution of matter, that is, relative to the fixed stars. E<strong>in</strong>ste<strong>in</strong>’s <strong>in</strong>terpretation of<br />

Mach’s statement was that the <strong>in</strong>ertial properties of a body is determ<strong>in</strong>ed by the presence of other bodies<br />

<strong>in</strong> the universe, and he named this concept “Mach’s Pr<strong>in</strong>ciple”.<br />

Equivalence pr<strong>in</strong>ciple:<br />

The equivalence pr<strong>in</strong>ciple comprises closely-related concepts deal<strong>in</strong>g with the equivalence of gravitational and<br />

<strong>in</strong>ertial mass. The weak equivalence pr<strong>in</strong>ciple states that the <strong>in</strong>ertial mass and gravitational mass of a<br />

body are identical, lead<strong>in</strong>g to acceleration that is <strong>in</strong>dependent of the nature of the body. Galileo demonstated<br />

this at the Lean<strong>in</strong>g Tower of Pisa. Recent measurements have shown that this weak equivalence pr<strong>in</strong>ciple<br />

is obeyed to a sensitivity of 5 × 10 −13 . E<strong>in</strong>ste<strong>in</strong>’s equivalence pr<strong>in</strong>ciple states that the outcome of<br />

any local non-gravitational experiment, <strong>in</strong> a freely fall<strong>in</strong>g laboratory, is <strong>in</strong>dependent of the velocity of the<br />

laboratory and its location <strong>in</strong> space-time. This pr<strong>in</strong>ciple implies that the result of local experiments must be<br />

<strong>in</strong>dependent of the velocity of the apparatus. E<strong>in</strong>ste<strong>in</strong>’s equivalence pr<strong>in</strong>ciple has been tested by search<strong>in</strong>g<br />

for variations of dimensionless fundamental constants such as the f<strong>in</strong>e structure constant. The strong<br />

equivalence pr<strong>in</strong>ciple comb<strong>in</strong>es the weak equivalence and E<strong>in</strong>ste<strong>in</strong> equivalence pr<strong>in</strong>ciples, and implies<br />

that the gravitational constant is constant everywhere <strong>in</strong> the universe. The strong equivalence pr<strong>in</strong>ciple<br />

suggests that gravity is geometrical <strong>in</strong> nature and does not <strong>in</strong>volve any fifth force <strong>in</strong> nature. Tests of the<br />

strong equivalence pr<strong>in</strong>ciple have <strong>in</strong>volved searches for variations <strong>in</strong> the gravitational constant and masses<br />

of fundamental particles throughout the life of the universe.<br />

Pr<strong>in</strong>ciple of covariance<br />

A physical law that is expressed <strong>in</strong> a covariant formulation has the same mathematical form <strong>in</strong> all coord<strong>in</strong>ate<br />

systems, and is usually expressed <strong>in</strong> terms of tensor fields. In the Special Theory of Relativity, the Lorentz,<br />

rotational, translational and reflection transformations between <strong>in</strong>ertial coord<strong>in</strong>ate frames are covariant. The<br />

covariant quantities are the 4-scalars, and 4-vectors <strong>in</strong> M<strong>in</strong>kowski space-time. E<strong>in</strong>ste<strong>in</strong> recognized that the<br />

pr<strong>in</strong>ciple of covariance, that is built <strong>in</strong>to the Special Theory of Relativity, should apply equally to accelerated<br />

relative motion <strong>in</strong> the General Theory of Relativity. He exploited tensor calculus to extend the Lorentz<br />

covariance to the more general local covariance <strong>in</strong> the General Theory of Relativity. The reduction locally<br />

of the general metric tensor to the M<strong>in</strong>kowski metric corresponds to free-fall<strong>in</strong>g motion, that is geodesic<br />

motion, and thus encompasses gravitation.


17.8. THE GENERAL THEORY OF RELATIVITY 481<br />

Pr<strong>in</strong>ciple of m<strong>in</strong>imal gravitational coupl<strong>in</strong>g<br />

The pr<strong>in</strong>ciple of m<strong>in</strong>imal gravitational coupl<strong>in</strong>g requires that the total Lagrangian for the field equations of<br />

general relativity consist of two additive parts, one part correspond<strong>in</strong>g to the free gravitational Lagrangian,<br />

and the other part to external source fields <strong>in</strong> curved space-time.<br />

Correspondence pr<strong>in</strong>ciple<br />

The Correspondence Pr<strong>in</strong>ciple states that the predictions of any new scientific theory must reduce to the<br />

predictions of well established earlier theories under circumstances for which the preced<strong>in</strong>g theory was known<br />

to be valid. The Correspondence Pr<strong>in</strong>ciple is an important concept used both <strong>in</strong> quantum mechanics and<br />

relativistic mechanics. E<strong>in</strong>ste<strong>in</strong>’s Special Theory of Relativity satisfies the Correspondence Pr<strong>in</strong>ciple because<br />

it reduces to classical mechanics <strong>in</strong> the limit of velocities small compared to the speed of light. The<br />

Correspondence Pr<strong>in</strong>ciple requires that the General Theory of Relativity reduce to the Special Theory of<br />

Relativity for <strong>in</strong>ertial frames, and should approximate Newton’s Theory of Gravitation <strong>in</strong> weak fields and at<br />

low velocities.<br />

17.8.2 E<strong>in</strong>ste<strong>in</strong>’s postulates for the General Theory of Relativity<br />

E<strong>in</strong>ste<strong>in</strong> realized that the Equivalence Pr<strong>in</strong>ciple relat<strong>in</strong>g the gravitational and <strong>in</strong>ertial masses implies that<br />

the constancy of the velocity of light <strong>in</strong> vacuum cannot hold <strong>in</strong> the presence of a gravitational field. That<br />

is, the M<strong>in</strong>kowskian l<strong>in</strong>e element must be replaced by a more general l<strong>in</strong>e element that takes gravity <strong>in</strong>to<br />

account. E<strong>in</strong>ste<strong>in</strong> proposed that the M<strong>in</strong>kowskian l<strong>in</strong>e element <strong>in</strong> four-dimensional space-time, be replaced<br />

by <strong>in</strong>troduc<strong>in</strong>g a four-dimensional Riemannian geometrical structure where space, time, and matter are<br />

comb<strong>in</strong>ed. As described by Lancos[La49], [Har03], [Mu08] this astonish<strong>in</strong>gly bold proposal implies that<br />

planetary motion is described as purely a geodesic phenomenon <strong>in</strong> a certa<strong>in</strong> four-space of Riemannian<br />

structure, where the geodesic is the equation of a curve on a manifold for any possible set of coord<strong>in</strong>ates.<br />

This implies that the concept of “gravitational force” is discarded, and planetary motion is a manifestation<br />

of a pure geodesic phenomenon for forceless motion <strong>in</strong> a four-dimensional Riemannian structure.<br />

Chapters 6 − 9 showed that the Lagrangian and Hamiltonian representations of variational pr<strong>in</strong>ciples are<br />

powerful approaches for determ<strong>in</strong><strong>in</strong>g the equation govern<strong>in</strong>g geodesic constra<strong>in</strong>ed motion that are <strong>in</strong>dependent<br />

of the chosen frame of reference as is also required by the General Theory of Relativity. Thus variational<br />

pr<strong>in</strong>ciples provide a theoretical representation for the General Theory of Relativity. The E<strong>in</strong>ste<strong>in</strong>-Hilbert<br />

action is def<strong>in</strong>ed as<br />

Z ∙<br />

¸<br />

1<br />

√−<br />

S =<br />

16 −4 R + L <br />

4 (17.96)<br />

where is E<strong>in</strong>ste<strong>in</strong>’s gravitational constant, R is the Ricci scalar, L accounts for matter fields, and is the<br />

determ<strong>in</strong>ant of the metric tensor.matrix. <strong>Variational</strong> pr<strong>in</strong>ciples applied to the E<strong>in</strong>ste<strong>in</strong>-Hilbert action lead to<br />

E<strong>in</strong>ste<strong>in</strong>’s sophisticated and advanced relativistic field equations of the General Theory of Relativity. Thus<br />

the variational approach unifies relativistic mechanics and classical field theories, such as mechanics and<br />

electromagnatism, which also were formulated <strong>in</strong> terms of least action. In relativistic mechanics, the use of<br />

action identifies the gravitational coupl<strong>in</strong>g of the metric to matter as well as identify<strong>in</strong>g conserved quantities<br />

and symmetries us<strong>in</strong>g Noether’s theorem. The E<strong>in</strong>ste<strong>in</strong>-Hilbert action expands the scope of variational<br />

pr<strong>in</strong>ciples to <strong>in</strong>clude general relativity illustrat<strong>in</strong>g the crucial role played by variational pr<strong>in</strong>ciples <strong>in</strong> physics.<br />

To summarize, the Special Theory of Relativity implies that the Newtonian concepts of absolute frame<br />

of reference and separation of space and time are <strong>in</strong>valid. The General Theory of Relativity goes beyond<br />

the Special Theory by imply<strong>in</strong>g that the gravitational force, and the resultant planetary motion, can be<br />

described as pure geodesic phenomena for forceless motion <strong>in</strong> a four-dimensional Riemannian structure.<br />

17.8.3 Experimental evidence <strong>in</strong> support of the General Theory of Relativity<br />

The follow<strong>in</strong>g experimental evidence <strong>in</strong> support of E<strong>in</strong>ste<strong>in</strong>’s Theory of General Relativity is compell<strong>in</strong>g.<br />

Kepler problem In 1915 E<strong>in</strong>ste<strong>in</strong> showed that relativistic mechanics expla<strong>in</strong>ed the anomalous precession<br />

of the perihelion of the planet mercury, that is, the axes of the elliptical Kepler orbit are observed to precess.


482 CHAPTER 17. RELATIVISTIC MECHANICS<br />

Deflection of light E<strong>in</strong>ste<strong>in</strong>’s prediction of the deflection of light <strong>in</strong> a gravitational field was confirmed by<br />

Edd<strong>in</strong>gton dur<strong>in</strong>g the solar eclipse of 29 May 1919 Pictures of stars <strong>in</strong> the region around the Sun showed that<br />

their apparent locations were slightly shifted because the light from the stars had been curved by pass<strong>in</strong>g<br />

close to the sun’s gravitational field.<br />

Gravitational lens<strong>in</strong>g The deflection of light by the gravitational attraction of a massive object situated<br />

between a distant star and the observer has resulted <strong>in</strong> the observation of multiple images of a distant quasar.<br />

Gravitational time dilation and frequency shift Processes occurr<strong>in</strong>g <strong>in</strong> a high gravitation field are<br />

slower than <strong>in</strong> a weak gravitational field; this is called gravitational time dilation. In addition, light climb<strong>in</strong>g<br />

out of a gravitational well is red shifted. The gravitational time dilation has been measured many times and<br />

the successful operation of the Global Position System provides an ongo<strong>in</strong>g validation. The gravitational<br />

redshifthasbeenconfirmed <strong>in</strong> the laboratory us<strong>in</strong>g the precise Mössbauer effect <strong>in</strong> nuclear physics. Tests<br />

<strong>in</strong> stronger gravitational fields are provided by studies of b<strong>in</strong>ary pulsars.<br />

Black holes When the mass to radius ratio of a massive object becomes sufficiently large, general relativity<br />

predicts formation of a black hole, which is a region of space from which neither light nor matter can escape.<br />

Supermassive black holes, with a mass that can be 10 6 − 10 9 solar masses, are thought to have played an<br />

important role <strong>in</strong> formation of the galaxies.<br />

Gravitational waves detection In 1916 E<strong>in</strong>ste<strong>in</strong> predicted the existence of gravitational waves on the<br />

basis of the theory of general relativity. The first implied detection of gravitational waves were made <strong>in</strong><br />

1976 by Hulse and Taylor who detected a decrease <strong>in</strong> the orbital period due to significant energy loss which<br />

presumably was associated with emission of gravity waves by the compact neutron star <strong>in</strong> the b<strong>in</strong>ary pulsar<br />

1913 + 16. The most compell<strong>in</strong>g direct evidence for observation of a gravitational wave was made<br />

on 15 September 2015 by the LIGO Laser Interferometer Gravitational-Wave Observatories. The waveform<br />

detected by the two LIGO observatories matched the predictions of General Relativity for gravitational waves<br />

emanat<strong>in</strong>g from the <strong>in</strong>ward spiral plus merger of a pair of black holes of around 36 and 29 solar masses,<br />

followed by the resultant b<strong>in</strong>ary black hole. The gravitational wave emitted by this cataclysmic merger<br />

reached Earth as a ripple <strong>in</strong> space-time that changed the length of the 4 LIGO arm by a thousandth of<br />

the width of the proton. The gravitational energy emitted was 30 +05<br />

−05 2 solar masses. A second observation<br />

of gravitational waves was made on 26 December 2015, and four similar observations were made dur<strong>in</strong>g<br />

2017. The detection of such m<strong>in</strong>iscule changes <strong>in</strong> space-time is a truly remarkable achievement. This direct<br />

detection of gravitational waves resulted <strong>in</strong> the award of the 2017 Nobel Prize to Ra<strong>in</strong>er Weiss, Barry<br />

Barish, and Kip Thorn. Gravitational wave detection has opened an excit<strong>in</strong>g and powerful new frontier <strong>in</strong><br />

astrophysics that could lead to excit<strong>in</strong>g new physics.<br />

17.9 Implications of relativistic theory to classical mechanics<br />

E<strong>in</strong>ste<strong>in</strong>’s theories of relativity have had an enormous impact on twentieth century physics and the philosophy<br />

of science. Relativistic mechanics is crucial to an understand<strong>in</strong>g of the physics of the atom, nucleus and the<br />

substructure of the nucleons, but the impacts are m<strong>in</strong>imal <strong>in</strong> everyday experience. The Special Theory of<br />

Relativity replaces Newton’s Laws of motion; i.e. Newton’s law is only an approximation applicable for low<br />

velocities. The General Theory of Relativity replaces Newton’s Law of Gravitation and provides a natural<br />

explanation of the equivalence pr<strong>in</strong>ciple. E<strong>in</strong>ste<strong>in</strong>’s theories of relativity imply a profound and fundamental<br />

change <strong>in</strong> the view of the separation of space, time, and mass, that contradicts the basic tenets that are<br />

the foundation of Newtonian mechanics. The Newtonian concepts of absolute frame of reference, plus the<br />

separation of space, time, and mass, are <strong>in</strong>valid at high velocities. Lagrangian and Hamiltonian variational<br />

approaches to classical mechanics provide the formalism necessary for handl<strong>in</strong>g relativistic mechanics. The<br />

present chapter has shown that logical extensions of Lagrangian and Hamiltonian mechanics lead to the<br />

relativistically-<strong>in</strong>variant extended Lagrangian and Hamiltonian formulations of mechanics which are adequate<br />

for handl<strong>in</strong>g one-body systems. However, major unsolved problems rema<strong>in</strong> apply<strong>in</strong>g these formulations to<br />

systems that have more than one body.


17.10. SUMMARY 483<br />

17.10 Summary<br />

Special theory of relativity: The Special Theory of Relativity is based on E<strong>in</strong>ste<strong>in</strong>’s postulates;<br />

1) The laws of nature are the same <strong>in</strong> all <strong>in</strong>ertial frames of reference.<br />

2) The velocity of light <strong>in</strong> vacuum is the same <strong>in</strong> all <strong>in</strong>ertial frames of reference.<br />

For a primed frame mov<strong>in</strong>g along the 1 axis with velocity E<strong>in</strong>ste<strong>in</strong>’s postulates imply the follow<strong>in</strong>g<br />

Lorentz transformations between the mov<strong>in</strong>g (primed) and stationary (unprimed) frames<br />

0 = ( − ) = ( 0 + 0 )<br />

0 = = 0<br />

0 = = 0<br />

0 = ¡ − <br />

2 ¢<br />

= <br />

³ ´<br />

0 + 0<br />

2<br />

1<br />

where the Lorentz factor ≡ <br />

1−( ) 2<br />

Lorentz transformations were used to illustrate Lorentz contraction, time dilation, and simultaneity. An<br />

elementary review was given of relativistic k<strong>in</strong>ematics <strong>in</strong>clud<strong>in</strong>g discussion of velocity transformation, l<strong>in</strong>ear<br />

momentum, center-of-momentum frame, forces and energy.<br />

Geometry of space-time: The concepts of four-dimensional space-time were <strong>in</strong>troduced. A discussion of<br />

four-vector scalar products <strong>in</strong>troduced the use of contravariant and covariant tensors plus the M<strong>in</strong>kowski metric<br />

where the scalar product was def<strong>in</strong>ed. The M<strong>in</strong>kowski representation of space time and the momentumenergy<br />

four vector also were <strong>in</strong>troduced.<br />

Lorentz-<strong>in</strong>variant formulation of Lagrangian mechanics: The Lorentz-<strong>in</strong>variant extended Lagrangian<br />

formalism, developed by Struckmeier[Str08], based on the parametric approach pioneered by<br />

Lanczos[La49], provides a viable Lorentz-<strong>in</strong>variant extension of conventional Lagrangian mechanics that<br />

is applicable for one-body motion <strong>in</strong> the realm of the Special Theory of Relativity.<br />

Lorentz-<strong>in</strong>variant formulation of Hamiltonian mechanics: The Lorentz-<strong>in</strong>variant extended Hamiltonian<br />

formalism, developed by Struckmeier based on the parametric approach pioneered by Lanczos, was<br />

<strong>in</strong>troduced. It provides a viable Lorentz-<strong>in</strong>variant extension of conventional Hamiltonian mechanics that is<br />

applicable for one-body motion <strong>in</strong> the realm of the Special Theory of Relativity. In particular, it was shown<br />

that the Lorentz-<strong>in</strong>variant extended Hamiltonian is conserved mak<strong>in</strong>g it ideally suited for solv<strong>in</strong>g complicated<br />

systems us<strong>in</strong>g Hamiltonian mechanics via use of the Poisson-bracket representation of Hamiltonian<br />

mechanics, canonical transformations, and the Hamilton-Jacobi techniques.<br />

The General Theory of Relativity: An elementary summary was given of the fundamental concepts<br />

of the General Theory of Relativity and the resultant unified description of the gravitational force plus<br />

planetary motion as geodesic motion <strong>in</strong> a four-dimensional Riemannian structure. <strong>Variational</strong> mechanics<br />

were shown to be ideally suited to applications of the General Theory of Relativity.<br />

Philosophical implications: Newton’s equations of motion, and his Law of Gravitation, that reigned<br />

supreme from 1687 to 1905, have been toppled from the throne by E<strong>in</strong>ste<strong>in</strong>’s theories of relativistic mechanics.<br />

By contrast, the complete <strong>in</strong>dependence to coord<strong>in</strong>ate frames <strong>in</strong> Lagrangian, and Hamiltonian formulations of<br />

classical mechanics, plus the underly<strong>in</strong>g Pr<strong>in</strong>ciple of Least Action, are equally valid <strong>in</strong> both the relativistic and<br />

non-relativistic regimes. As a consequence, relativistic Lagrangian and Hamiltonian formulations underlie<br />

much of modern physics, especially quantum physics, which expla<strong>in</strong>s why relativistic mechanics plays such<br />

an important role <strong>in</strong> classical dynamics.


484 CHAPTER 17. RELATIVISTIC MECHANICS<br />

Problems<br />

1. A relativistic snake of proper length 100 istravell<strong>in</strong>gtotherightacrossabutcher’stableat =06. You<br />

hold two meat cleavers, one <strong>in</strong> each hand which are 100 apart. You strike the table simultaneously with<br />

both cleavers at the moment when the left cleaver lands just beh<strong>in</strong>d the tail of the snake. You rationalize that<br />

s<strong>in</strong>ce the snake is mov<strong>in</strong>g with =06 then the length of the snake is Lorentz contracted by the factor = 5 4<br />

and thus the Lorentz-contracted length of the snake is 80 and thus will not be harmed. However, the snake<br />

reasons that relative to it the cleavers are mov<strong>in</strong>g at =06 and thus are only 80 apart when they strike<br />

the 100 long snake and thus it will be severed. Use the Lorentz transformation to resolve this paradox.<br />

2. Expla<strong>in</strong> what is meant by the follow<strong>in</strong>g statement: “Lorentz transformations are orthogonal transformations<br />

<strong>in</strong> M<strong>in</strong>kowski space.”<br />

3. Which of the follow<strong>in</strong>g are <strong>in</strong>variant quantities <strong>in</strong> space-time?<br />

(a) Energy<br />

(b) Momentum<br />

(c) Mass<br />

(d) Force<br />

(e) Charge<br />

(f) The length of a vector<br />

(g) The length of a four-vector<br />

4. What does it mean for two events to have a spacelike <strong>in</strong>terval? What does it mean for them to have a timelike<br />

<strong>in</strong>terval? Draw a picture to support your answer. In which case can events be causally connected?<br />

5. A supply rocket flies past two markers on the Space Station that are 50 apart <strong>in</strong> a time of 02 as measured<br />

by an observer on the Space station.<br />

(a) What is the separation of the two markers as seen by the pilot rid<strong>in</strong>g <strong>in</strong> the supply rocket?<br />

(b) What is the elapsed time as measured by the pilot <strong>in</strong> the supply rocket?<br />

(c) What are the speeds calculated by the observer <strong>in</strong> the Space Station and the pilot of the supply rocket?<br />

6. The Compton effect <strong>in</strong>volves a photon of <strong>in</strong>cident energy be<strong>in</strong>g scattered by an electron of mass which<br />

<strong>in</strong>itially is stationary. The photon scattered at an angle with respect to the <strong>in</strong>cident photon has a f<strong>in</strong>al energy<br />

. Us<strong>in</strong>g the special theory of relativity derive a formula that related and to .<br />

7. Pair creation <strong>in</strong>volves production of an electron-positron pair by a photon. Show that such a process is<br />

impossible unless some other body, such as a nucleus, is <strong>in</strong>volved. Suppose that the nucleus has a mass <br />

and the electron mass . What is the m<strong>in</strong>imum energy that the photon must have <strong>in</strong> order to produce an<br />

electron-positron pair?<br />

8. A meson of rest energy 494 decays <strong>in</strong>to a meson of rest energy 106 and a neutr<strong>in</strong>o of zero<br />

rest energy. F<strong>in</strong>d the k<strong>in</strong>etic energies of the mesonandtheneutr<strong>in</strong>o<strong>in</strong>towhichthe meson decays while<br />

at rest.


Chapter 18<br />

The transition to quantum physics<br />

18.1 Introduction<br />

<strong>Classical</strong> mechanics, <strong>in</strong>clud<strong>in</strong>g extensions to relativistic velocities, embrace an unusually broad range of topics<br />

rang<strong>in</strong>g from astrophysics to nuclear and particle physics, from one-body to many-body statistical mechanics.<br />

It is <strong>in</strong>terest<strong>in</strong>g to discuss the role of classical mechanics <strong>in</strong> the development of quantum mechanics which<br />

plays a crucial role <strong>in</strong> physics. A valid question is “why discuss quantum mechanics <strong>in</strong> a classical mechanics<br />

course?”. The answer is that quantum mechanics supersedes classical mechanics as the fundamental theory<br />

of mechanics. <strong>Classical</strong> mechanics is an approximation applicable for situations where quantization is<br />

unimportant. Thus there must be a correspondence pr<strong>in</strong>ciple that relates quantum mechanics to classical<br />

mechanics, analogous to the relation between relativistic and non-relativistic mechanics. It is illum<strong>in</strong>at<strong>in</strong>g to<br />

study the role played by the Hamiltonian formulation of classical mechanics <strong>in</strong> the development of quantal<br />

theory and statistical mechanics. The Hamiltonian formulation is expressed <strong>in</strong> terms of the phase-space<br />

variables q p for which there are well-established rules for transform<strong>in</strong>g to quantal l<strong>in</strong>ear operators.<br />

18.2 Brief summary of the orig<strong>in</strong>s of quantum theory<br />

The last decade of the 19 century saw the culm<strong>in</strong>ation of classical physics. By 1900 scientists thought<br />

that the basic laws of mechanics, electromagnetism, and statistical mechanics were understood and worried<br />

that future physics would be reduced to confirm<strong>in</strong>g theories to the fifth decimal place, with few major new<br />

discoveries to be made. However, technical developments such as photography, vacuum pumps, <strong>in</strong>duction<br />

coil, etc., led to important discoveries that revolutionized physics and toppled classical mechanics from its<br />

throne at the beg<strong>in</strong>n<strong>in</strong>g of the 20 century. Table 181 summarizes some of the major milestones lead<strong>in</strong>g<br />

up to the development of quantum mechanics.<br />

Max Planck searched for an explanation of the spectral shape of the black-body electromagnetic radiation.<br />

He found an <strong>in</strong>terpolation between two conflict<strong>in</strong>g theories, one that reproduced the short wavelength<br />

behavior, and the other the long wavelength behavior. Planck’s <strong>in</strong>terpolation required assum<strong>in</strong>g that electromagnetic<br />

radiation was not emitted with a cont<strong>in</strong>uous range of energies, but that electromagnetic radiation<br />

is emitted <strong>in</strong> discrete bundles of energy called quanta. In December 1900 he presented his theory which<br />

reproduced precisely the measured black body spectral distribution by assum<strong>in</strong>g that the energy carried by<br />

a s<strong>in</strong>gle quantum must be an <strong>in</strong>teger multiple of :<br />

= = <br />

<br />

(18.1)<br />

where is the frequency of the electromagnetic radiation and Planck’s constant, =662610 −34 · was<br />

the best fit parameter of the <strong>in</strong>terpolation. That is, Planck assumed that energy comes <strong>in</strong> discrete bundles<br />

of energy equal to which are called quanta. By mak<strong>in</strong>g this extreme assumption, <strong>in</strong> an act of desperation,<br />

Planck was able to reproduce the experimental black body radiation spectrum. The assumption that energy<br />

was exchanged <strong>in</strong> bundles h<strong>in</strong>ted that the classical laws of physics were <strong>in</strong>adequate <strong>in</strong> the microscopic<br />

doma<strong>in</strong>. The older generation physicists <strong>in</strong>itially refused to believe Planck’s hypothesis which underlies<br />

485


486 CHAPTER 18. THE TRANSITION TO QUANTUM PHYSICS<br />

quantum theory. It was the new generation physicists, like E<strong>in</strong>ste<strong>in</strong>, Bohr, Heisenberg, Born, Schröd<strong>in</strong>ger,<br />

and Dirac, who developed Planck’s hypothesis lead<strong>in</strong>g to the revolutionary quantum theory.<br />

In 1905, E<strong>in</strong>ste<strong>in</strong> predicted the existence of the photon, derived the theory of specific heat,aswell<br />

as deriv<strong>in</strong>g the Theory of Special Relativity. It is remarkable to realize that he developed these three<br />

revolutionary theories <strong>in</strong> one year, when he was only 26 years old. E<strong>in</strong>ste<strong>in</strong> uncovered an <strong>in</strong>consistency <strong>in</strong><br />

Planck’s derivation of the black body spectral distribution<strong>in</strong>thatitassumedthestatisticalpartoftheenergy<br />

is quantized, whereas the electromagnetic radiation assumed Maxwell’s equations with oscillator energies<br />

be<strong>in</strong>g cont<strong>in</strong>uous. Planck demanded that light of frequency be packaged <strong>in</strong> quanta whose energies were<br />

multiples of , but Planck never thought that light would have particle-like behavior. Newton believed that<br />

light <strong>in</strong>volved corpuscles, and Hamilton developed the Hamilton-Jacobi theory seek<strong>in</strong>g to describe light <strong>in</strong><br />

terms of the corpuscle theory. However, Maxwell had conv<strong>in</strong>ced physicists that light was a wave phenomena;<br />

<strong>in</strong>terference plus diffraction effects were conv<strong>in</strong>c<strong>in</strong>g manifestations of the wave-like properties of light. In<br />

order to reproduce Planck’s prediction, E<strong>in</strong>ste<strong>in</strong> had to treat black-body radiation as if it consisted of a gas<br />

of photons, each photon hav<strong>in</strong>g energy = . This was a revolutionary concept that returned to Newton’s<br />

corpuscle theory of light. E<strong>in</strong>ste<strong>in</strong> realized that there were direct tests of his photon hypothesis, one of which<br />

is the photo-electric effect. Accord<strong>in</strong>g to E<strong>in</strong>ste<strong>in</strong>, each photon has an energy = , <strong>in</strong> contrast to the<br />

classical case where the energy of the photoelectron depends on the <strong>in</strong>tensity of the light. E<strong>in</strong>ste<strong>in</strong> predicted<br />

that the ejected electron will have a k<strong>in</strong>etic energy<br />

= − (18.2)<br />

where is the work function which is the energy needed to remove an electron from a solid.<br />

Many older scientists, <strong>in</strong>clud<strong>in</strong>g Planck, accepted E<strong>in</strong>ste<strong>in</strong>’s theory of relativity but were skeptical of the<br />

photon concept, even after E<strong>in</strong>ste<strong>in</strong>’s photon concept was v<strong>in</strong>dicated <strong>in</strong> 1915 by Millikan who showed that,<br />

as predicted, the energy of the ejected photoelectron depended on the frequency, and not <strong>in</strong>tensity, of the<br />

light. In 1923 Compton’s demonstrated that electromagnetic radiation scattered by free electrons obeyed<br />

simple two-body scatter<strong>in</strong>g laws which f<strong>in</strong>ally conv<strong>in</strong>ced the many skeptics of the existence of the photon.<br />

Table 181: Chronology of the development of quantum mechanics<br />

Date Author Development<br />

1887 Hertz Discovered the photo-electric effect<br />

1895 Röntgen Discovered x-rays<br />

1896 Becquerel Discovered radioactivity<br />

1897 J.J. Thomson Discovered the first fundamental particle, the electron<br />

1898 Pierre & Marie Curie Showed that thorium is radioactive which founded nuclear physics<br />

1900 Planck Quantization = expla<strong>in</strong>ed the black-body spectrum<br />

1905 E<strong>in</strong>ste<strong>in</strong> Theory of special relativity<br />

1905 E<strong>in</strong>ste<strong>in</strong> Predicted the existence of the photon<br />

1906 E<strong>in</strong>ste<strong>in</strong> Used Planck’s constant to expla<strong>in</strong> specific heats of solids<br />

1909 Millikan The oil drop experiment measured the charge on the electron<br />

1911 Rutherford Discovered the atomic nucleus with radius 10 −15 <br />

1912 Bohr Bohr model of the atom expla<strong>in</strong>ed the quantized states of hydrogen<br />

1914 Moseley X-ray spectra determ<strong>in</strong>ed the atomic number of the elements.<br />

1915 Millikan Used the photo-electric effect to confirm the photon hypothesis.<br />

1915 Wilson-Sommerfeld Proposed quantization of the action-angle <strong>in</strong>tegral<br />

1921 Stern-Gerlach Observed space quantization <strong>in</strong> non-uniform magnetic field<br />

1923 Compton Compton scatter<strong>in</strong>g of x-rays confirmed the photon hypothesis<br />

1924 de Broglie Postulated wave-particle duality for matter and EM waves<br />

1924 Bohr Explicit statement of the correspondence pr<strong>in</strong>ciple<br />

1925 Pauli Postulated the exclusion pr<strong>in</strong>ciple<br />

1925 Goudsmit-Uhlenbeck Postulated the sp<strong>in</strong> of the electron of = 1 2 h<br />

1925 Heisenberg Matrix mechanics representation of quantum theory<br />

1925 Dirac Related Poisson brackets and commutation relations<br />

1926 Schröd<strong>in</strong>ger Wave mechanics<br />

1927 G.P. Thomson/Davisson Electron diffraction proved wave nature of electron<br />

1928 Dirac Developed the Dirac relativistic wave equation


18.2. BRIEF SUMMARY OF THE ORIGINS OF QUANTUM THEORY 487<br />

18.2.1 Bohr model of the atom<br />

The Rutherford scatter<strong>in</strong>g experiment, performed at Manchester <strong>in</strong> 1911, discovered that the Au atom<br />

comprised a positively charge nucleus of radius ≈ 10 −14 which is much smaller than the 135 × 10 −10 <br />

radius of the Au atom. Stimulated by this discovery, Niels Bohr jo<strong>in</strong>ed Rutherford at Manchester <strong>in</strong> 1912<br />

where he developed the Bohr model of the atom. This theory was remarkably successful <strong>in</strong> spite of hav<strong>in</strong>g<br />

serious <strong>in</strong>consistencies and deficiencies. Bohr’s model assumptions were:<br />

1) Electromagnetic radiation is quantized with = <br />

2) Electromagnetic radiation exhibits behavior characteristic of the emission of photons with energy<br />

= and momentum = <br />

<br />

. That is, it exhibits both wave-like and particle-like behavior.<br />

3) Electrons are <strong>in</strong> stationary orbits that do not radiate, which contradicts the predictions of classical<br />

electromagnetism.<br />

<br />

4) The orbits are quantized such that the electron angular momentum is an <strong>in</strong>teger multiple of<br />

2 = ~<br />

5) Atomic electromagnetic radiation is emitted with photon energy equal to the difference <strong>in</strong> b<strong>in</strong>d<strong>in</strong>g<br />

energy between the two atomic levels <strong>in</strong>volved. = 1 − 2<br />

The first two assumptions are due to Planck and E<strong>in</strong>ste<strong>in</strong>, while the last three were made by Niels Bohr.<br />

The deficiencies of the Bohr model were the philosophical problems of violat<strong>in</strong>g the tenets of classical<br />

physics <strong>in</strong> expla<strong>in</strong><strong>in</strong>g hydrogen-like atoms, that is, the theory was prescriptive, not deductive. The Bohr<br />

model was based implicitly on the assumption that quantum theory conta<strong>in</strong>s classical mechanics as a limit<strong>in</strong>g<br />

case. Bohr explicitly stated this assumption which he called the correspondence pr<strong>in</strong>ciple, and which<br />

played a pivotal role <strong>in</strong> the development of the older quantum theory. In 1924 Bohr justified the <strong>in</strong>consistencies<br />

of the old quantum theory by writ<strong>in</strong>g “As frequently emphasized, these pr<strong>in</strong>ciples, although they<br />

are formulated by the help of classical conceptions, are to be regarded purely as laws of quantum theory,<br />

which give us, not withstand<strong>in</strong>g the formal nature of quantum theory, a hope <strong>in</strong> the future of a consistent<br />

theory, which at the same time reproduces the characteristic features of quantum theory, important for its<br />

applicability, and, nevertheless, can be regarded as a rational generalization of classical electrodynamics.”<br />

The old quantum theory was remarkably successful <strong>in</strong> reproduc<strong>in</strong>g the black-body spectrum, specificheats<br />

of solids, the hydrogen atom, and the periodic table of the elements. Unfortunately, from a methodological<br />

po<strong>in</strong>t of view, the theory was a hodgepodge of hypotheses, pr<strong>in</strong>ciples, theorems, and computational recipes,<br />

rather than a logical consistent theory. Every problem was first solved <strong>in</strong> terms of classical mechanics,<br />

and then would pass through a mysterious quantization procedure <strong>in</strong>volv<strong>in</strong>g the correspondence pr<strong>in</strong>ciple.<br />

Although built on the foundation of classical mechanics, it required Bohr’s hypotheses which violated the<br />

laws of classical mechanics and predictions of Maxwell’s equations.<br />

18.2.2 Quantization<br />

By 1912 Planck, and others, had abandoned the concept that quantum theory was a branch of classical<br />

mechanics, and were search<strong>in</strong>g to see if classical mechanics was a special case of a more general quantum<br />

physics, or quantum physics was a science altogether outside of classical mechanics. Also they were try<strong>in</strong>g<br />

to f<strong>in</strong>d a consistent and rational reason for quantization to replace the ad hoc assumption of Bohr.<br />

In 1912 Sommerfeld proposed that, <strong>in</strong> every elementary process, the atom ga<strong>in</strong>s or loses a def<strong>in</strong>ite amount<br />

of action between times 0 and of<br />

Z <br />

= ( 0 ) 0 (18.3)<br />

0<br />

where is the quantal analogue of the classical action function It has been shown that the classical pr<strong>in</strong>ciple<br />

of least action states that the action function is stationary for small variations of the trajectory. In 1915<br />

Wilson and Sommerfeld recognized that the quantization of angular momentum could be expressed <strong>in</strong> terms<br />

of the action-angle <strong>in</strong>tegral, that is equation 15116. They postulated that, for every coord<strong>in</strong>ate, the actionangle<br />

variable is quantized<br />

I<br />

= (18.4)<br />

where the action-angle variable <strong>in</strong>tegral is over one complete period of the motion. That is, they postulated<br />

that Hamilton’s phase space is quantized, but the microscopic granularity is such that the quantization is<br />

only manifest for atomic-sized doma<strong>in</strong>s. That is, is a small <strong>in</strong>teger for atomic systems <strong>in</strong> contrast to<br />

≈ 10 64 for the Earth-Sun two-body system.


488 CHAPTER 18. THE TRANSITION TO QUANTUM PHYSICS<br />

Sommerfeld recognized that quantization of more than one degree of freedom is needed to obta<strong>in</strong> more<br />

accurate description of the hydrogen atom. Sommerfeld reproduced the experimental data by assum<strong>in</strong>g<br />

quantization of the three degrees of freedom,<br />

I<br />

I<br />

I<br />

= 1 <br />

= 2 <br />

= 3 (18.5)<br />

and solv<strong>in</strong>g Hamilton-Jacobi theory by separation of variables. In 1916 the Bohr-Sommerfeld model solved<br />

the classical orbits for the hydrogen atom, <strong>in</strong>clud<strong>in</strong>g relativistic corrections as described <strong>in</strong> example 177.<br />

This reproduced f<strong>in</strong>e structure observed <strong>in</strong> the optical spectra of hydrogen. The use of the canonical transformation<br />

to action-angle variables proved to be the ideal approach for solv<strong>in</strong>g many such problems <strong>in</strong><br />

quantum mechanics. In 1921 Stern and Gerlach demonstrated space quantization by observ<strong>in</strong>g the splitt<strong>in</strong>g<br />

of atomic beams deflected by non-uniform magnetic fields. This result was a major triumph for quantum<br />

theory. Sommerfeld declared that “With their bold experimental method, Stern and Gerlach demonstrated<br />

not only the existence of space quantization, they also proved the atomic nature of the magnetic moment,<br />

its quantum-theoretic orig<strong>in</strong>, and its relation to the atomic structure of electricity.”<br />

In 1925 Pauli’s Exclusion Pr<strong>in</strong>ciple proposed that no more than one electron can have identical quantum<br />

numbers and that the atomic electronic state is specified by four quantum numbers. Two students, Goudsmit<br />

and Uhlenbeck suggested that a fourth two-valued quantum number was the electron sp<strong>in</strong> of ± 2 . This<br />

provided a plausible explanation for the structure of multi-electron atoms.<br />

18.2.3 Wave-particle duality<br />

In his 1924 doctoral thesis, Pr<strong>in</strong>ce Louis de Broglie proposed the hypothesis of wave-particle duality which<br />

was a pivotal development <strong>in</strong> quantum theory. de Broglie used the classical concept of a matter wavepacket,<br />

analogous to classical wave packets discussed <strong>in</strong> chapter 311. He assumed that both the group and signal<br />

velocities of a matter wave packet must equal the velocity of the correspond<strong>in</strong>g particle. By analogy with<br />

E<strong>in</strong>ste<strong>in</strong>’s relation for the photon, and us<strong>in</strong>g the Theory of Special Relativity, de Broglie assumed that<br />

~ = =<br />

2<br />

q ¡1<br />

−<br />

2<br />

2 ¢ (18.6)<br />

The group velocity is required to equal the velocity of the mass <br />

µ µ µ <br />

<br />

= =<br />

= (18.7)<br />

<br />

This gives<br />

<br />

= 1 <br />

µ <br />

³ <br />

´ µ1<br />

− 3<br />

= − 2 2<br />

~ 2 (18.8)<br />

Integration of this equation assum<strong>in</strong>g that =0when =0,thengives<br />

v<br />

~k = q ¡1 ¢ = p (18.9)<br />

−<br />

v·v<br />

2<br />

This relation, derived by de Broglie, is required to ensure that the particle travels at the group velocity<br />

of the wave packet characteriz<strong>in</strong>g the particle. Note that although the relations used to characterize the<br />

matter waves are purely classical, the physical content of such waves is beyond classical physics. In 1927 C.<br />

Davisson and G.P. Thomson <strong>in</strong>dependently observed electron diffraction confirm<strong>in</strong>g wave/particle duality for<br />

the electron. Ironically, J.J. Thomson discovered that the electron was a particle, whereas his son attributed<br />

it to an electron wave.<br />

Heisenberg developed the modern matrix formulation of quantum theory <strong>in</strong> 1925; hewas24 years old<br />

at the time. A few months later Schröd<strong>in</strong>ger’s developed wave mechanics based on de Broglie’s concept of<br />

wave-particle duality. The matrix mechanics, and wave mechanics, quantum theories are radically different.<br />

Heisenberg’s algebraic approach employs non-commut<strong>in</strong>g quantities and unfamiliar mathematical techniques<br />

that emphasized the discreteness characteristic of the corpuscle aspect. In contrast, Schröd<strong>in</strong>ger used the<br />

familiar analytical approach that is an extension of classical laws of motion and waves which stressed the<br />

element of cont<strong>in</strong>uity.


18.3. HAMILTONIAN IN QUANTUM THEORY 489<br />

18.3 Hamiltonian <strong>in</strong> quantum theory<br />

18.3.1 Heisenberg’s matrix-mechanics representation<br />

The algebraic Heisenberg representation of quantum theory is analogous to the algebraic Hamiltonian representation<br />

of classical mechanics, and shows best how quantum theory evolved from, and is related to,<br />

classical mechanics. Heisenberg decided to ignore the prevail<strong>in</strong>g conceptual theories, such as classical mechanics,<br />

and based his quantum theory on observables. This approach was <strong>in</strong>fluenced by the success of<br />

Bohr’s older quantum theory and E<strong>in</strong>ste<strong>in</strong>’s theory of relativity. He abandoned the classical notions that<br />

the canonical variables can be measured directly and simultaneously. <strong>Second</strong>ly he wished to absorb the<br />

correspondence pr<strong>in</strong>ciple directly <strong>in</strong>to the theory <strong>in</strong>stead of it be<strong>in</strong>g an ad hoc procedure tailored to each application.<br />

Heisenberg considered the Fourier decomposition of transition amplitudes between discrete states<br />

and found that the product of the conjugate variables do not commute. Heisenberg derived, for the first<br />

time, the correct energy levels of the one-dimensional harmonic oscillator as = ~( + 1 2<br />

) which was a<br />

significant achievement. Born recognized that Heisenberg’s strange multiplication and commutation rules for<br />

two variables, corresponded to matrix algebra. Prior to 1925 matrix algebra was an obscure branch of pure<br />

mathematics not known or used by the physics community. Heisenberg, Born, and the young mathematician<br />

Jordan, developed the commutation rules of matrix mechanics. Heisenberg’s approach represents the<br />

classical position and momentum coord<strong>in</strong>ates by matrices q and p, with correspond<strong>in</strong>g matrix elements<br />

and . Born showed that the trace of the matrix<br />

(pq) =p ˙q− (18.10)<br />

gives the Hamiltonian function (p q) of the matrices q and p which leads to Hamilton’s canonical equations<br />

˙q= <br />

p<br />

ṗ=− <br />

q<br />

(18.11)<br />

Heisenberg and Born also showed that the commutator of q p equals<br />

− = ~ (18.12)<br />

− = 0<br />

− = 0<br />

Born realized that equation (1812) is the only fundamental equation for <strong>in</strong>troduc<strong>in</strong>g ~ <strong>in</strong>to the theory <strong>in</strong> a<br />

logical and consistent way.<br />

Chapter 1524 discussed the formal correspondence between the Poisson bracket, def<strong>in</strong>ed<strong>in</strong>chapter153,<br />

and the commutator <strong>in</strong> classical mechanics. It was shown that the commutator of two functions equals a<br />

constant multiplicative factor times the correspond<strong>in</strong>g Poisson Bracket. That is<br />

( − )= [ ] (18.13)<br />

where the multiplicative factor is a number <strong>in</strong>dependent of , and the commutator.<br />

In 1925, Paul Dirac, a 23-year old graduate student at Bristol, recognized the crucial importance of<br />

the above correspondence between the commutator and the Poisson Bracket of two functions, to relat<strong>in</strong>g<br />

classical mechanics and quantum mechanics. Dirac noted that if the constant is assigned the value = ~,<br />

then equation 1813 directly relates Heisenberg’s commutation relations between the fundamental canonical<br />

variables ( ) to the correspond<strong>in</strong>g classical Poisson Bracket [ ].Thatis,<br />

− = ~ [ ]=~ (18.14)<br />

− = ~ [ ]=0 (18.15)<br />

− = ~ [ ]=0 (18.16)<br />

Dirac recognized that the correspondence between the classical Poisson bracket, and quantum commutator,<br />

given by equation (1813) provides a logical and consistent way that builds quantization directly <strong>in</strong>to<br />

the theory, rather than us<strong>in</strong>g an ad-hoc, case-dependent, hypothesis as used by the older quantum theory of


490 CHAPTER 18. THE TRANSITION TO QUANTUM PHYSICS<br />

Bohr. The basis of Dirac’s quantization pr<strong>in</strong>ciple, <strong>in</strong>volves replac<strong>in</strong>g the classical Poisson Bracket, [ ]<br />

by the commutator,<br />

1<br />

( − ).Thatis,<br />

[ ]=⇒ 1<br />

~ ( − ) (18.17)<br />

Hamilton’s canonical equations, as <strong>in</strong>troduced <strong>in</strong> chapter 15, are only applicable to classical mechanics<br />

s<strong>in</strong>ce they assume that the exact position and conjugate momentum can be specified both exactly and<br />

simultaneously which contradicts the Heisenberg’s Uncerta<strong>in</strong>ty Pr<strong>in</strong>ciple. In contrast, the Poisson bracket<br />

generalization of Hamilton’s equations allows for non-commut<strong>in</strong>g variables plus the correspond<strong>in</strong>g uncerta<strong>in</strong>ty<br />

pr<strong>in</strong>ciple. That is, the transformation from classical mechanics to quantum mechanics can be accomplished<br />

simply by replac<strong>in</strong>g the classical Poisson Bracket by the quantum commutator, as proposed by Dirac. The<br />

formal analogy between classical Hamiltonian mechanics, and the Heisenberg representation of quantum<br />

mechanics is strik<strong>in</strong>gly apparent us<strong>in</strong>g the correspondence between the Poisson Bracket representation of<br />

Hamiltonian mechanics and Heisenberg’s matrix mechanics.<br />

The direct relation between the quantum commutator, and the correspond<strong>in</strong>g classical Poisson Bracket,<br />

applies to many observables. For example, the quantum analogs of Hamilton’s equations of motion are<br />

given by use of Hamilton’s equations of motion, 1553 1556 and replac<strong>in</strong>g each Poisson Bracket by the<br />

correspond<strong>in</strong>g commutator. That is<br />

<br />

<br />

<br />

<br />

= <br />

<br />

=[ ]= 1<br />

~ ( − ) (18.18)<br />

= − <br />

<br />

=[ ]= 1<br />

~ ( − ) (18.19)<br />

Chapter 1525 discussed the time dependence of observables <strong>in</strong> Hamiltonian mechanics. Equation 1545<br />

gave the total time derivative of any observable to be<br />

<br />

= +[ ] (18.20)<br />

<br />

Equation 1817 can be used to replace the Poisson Bracket by the quantum commutator, which gives the<br />

correspond<strong>in</strong>g time dependence of observables <strong>in</strong> quantum physics.<br />

<br />

= <br />

+ 1 ( − ) (18.21)<br />

~<br />

In quantum mechanics, equation 1821 is called the Heisenberg equation. Note that if the observable is<br />

chosen to be a fundamental canonical variable, then <br />

<br />

=0= <br />

<br />

and equation 1520 reduces to Hamilton’s<br />

equations 1818 and 1819.<br />

The analogies between classical mechanics and quantum mechanics extend further. For example, if is<br />

a constant of motion, that is <br />

<br />

=0 then Heisenberg’s equation of motion gives<br />

Moreover, if is not an explicit function of time, then<br />

<br />

+ 1 ( − )=0 (18.22)<br />

~<br />

0= 1 ( − ) (18.23)<br />

~<br />

That is, the transition to quantum physics shows that, if is a constant of motion, and is not explicitly<br />

time dependent, then commutes with the Hamiltonian .<br />

The above discussion has illustrated the close and beautiful correspondence between the Poisson Bracket<br />

representation of classical Hamiltonian mechanics, and the Heisenberg representation of quantum mechanics.<br />

Dirac provided the elegant and simple correspondence pr<strong>in</strong>ciple connect<strong>in</strong>g the Poisson bracket representation<br />

of classical Hamiltonian mechanics, to the Heisenberg representation of quantum mechanics.


18.3. HAMILTONIAN IN QUANTUM THEORY 491<br />

18.3.2 Schröd<strong>in</strong>ger’s wave-mechanics representation<br />

The wave mechanics formulation of quantum mechanics, by the Austrian theorist Schröd<strong>in</strong>ger, was built on<br />

the wave-particle duality concept that was proposed <strong>in</strong> 1924 by Louis de Broglie. Schröd<strong>in</strong>ger developed<br />

his wave mechanics representation of quantum physics a year after the development of matrix mechanics<br />

by Heisenberg and Born. The Schröd<strong>in</strong>ger wave equation is based on the non-relativistic Hamilton-Jacobi<br />

representation of a wave equation, melded with the operator formalism of Born and Wiener. The 39-year old<br />

Schröd<strong>in</strong>ger was an expert <strong>in</strong> classical mechanics and wave theory, which was <strong>in</strong>valuable when he developed<br />

the important Schröd<strong>in</strong>ger equation. As mentioned <strong>in</strong> chapter 1544, the Hamilton-Jacobi theory is a<br />

formalism of classical mechanics that allows the motion of a particle to be represented by a wave. That is,<br />

the wavefronts are surfaces of constant action and the particle momenta are normal to these constantaction<br />

surfaces, that is, p = ∇. The wave-particle duality of Hamilton-Jacobi theory is a natural way to<br />

handle the wave-particle duality proposed by de Broglie.<br />

Consider the classical Hamilton-Jacobi equation for one body, given by 1820<br />

<br />

+ (q ∇)=0 (18.24)<br />

<br />

If the Hamiltonian is time <strong>in</strong>dependent, then equation 1590 gives that<br />

<br />

= −(q p)=− (α) (18.25)<br />

<br />

The <strong>in</strong>tegration of the time dependence is trivial, and thus the action <strong>in</strong>tegral for a time-<strong>in</strong>dependent Hamiltonian<br />

is<br />

(q α) = (q α) − (α) (18.26)<br />

A formal transformation gives<br />

= − <br />

p = ∇ (18.27)<br />

<br />

Consider that the classical time-<strong>in</strong>dependent Hamiltonian, for motion of a s<strong>in</strong>gle particle, is represented<br />

by the Hamilton-Jacobi equation.<br />

= p2<br />

+ () =−<br />

(18.28)<br />

2 <br />

Substitute for p leads to the classical Hamilton-Jacobi relation <strong>in</strong> terms of the action <br />

1<br />

(∇ · ∇)+() =−<br />

2 <br />

(18.29)<br />

By analogy with the Hamilton-Jacobi equation, Schröd<strong>in</strong>ger proposed the quantum operator equation<br />

~ <br />

= ˆ (18.30)<br />

where ˆ is an operator given by<br />

ˆ = − ~2<br />

2 ∇2 + () (18.31)<br />

In 1926 Max Born and Norbert Wiener <strong>in</strong>troduced the operator formalism <strong>in</strong>to matrix mechanics for prediction<br />

of observables and this has become an <strong>in</strong>tegral part of quantum theory. In the operator formalism, the<br />

observables are represented by operators that project the correspond<strong>in</strong>g observable from the wavefunction.<br />

That is, the quantum operator formalism for the assumed momentum and energy operators, that operate<br />

on the wavefunction , are<br />

= ~ <br />

= − ~ <br />

(18.32)<br />

<br />

<br />

Formal transformations of p and <strong>in</strong> the Hamiltonian (1826) leads to the time-<strong>in</strong>dependent Schröd<strong>in</strong>ger<br />

equation<br />

− ~2 2 <br />

+ () = (18.33)<br />

2 2


492 CHAPTER 18. THE TRANSITION TO QUANTUM PHYSICS<br />

Assume that the wavefunction is of the form<br />

= (18.34)<br />

where the action gives the phase of the wavefront, and the amplitude of the wave, as described <strong>in</strong><br />

chapter 1544. The time dependence, that characterizes the motion of the wavefront, is conta<strong>in</strong>ed <strong>in</strong> the<br />

time dependence of This form for the wavefunction has the advantage that the wavefunction frequently<br />

factors <strong>in</strong>to a product of terms, e.g. = ()Θ()Φ() which corresponds to a summation of the exponents<br />

= + + − . This summation form is exploited by separation of the variables, as discussed <strong>in</strong><br />

chapter 1543.<br />

Insert (1834) <strong>in</strong>to equation (1833) plus us<strong>in</strong>g the fact that<br />

2 <br />

2<br />

= <br />

µ <br />

<br />

<br />

<br />

= <br />

µ <br />

~ <br />

= − 1 µ 2 <br />

~ 2 + ~ 2 <br />

2 (18.35)<br />

leads to<br />

− <br />

= 1<br />

~<br />

(∇ · ∇)+() −<br />

2 2 ∇2 = (18.36)<br />

Note that if Planck’s constant ~ =0 then the imag<strong>in</strong>ary term <strong>in</strong> equation (1836) is zero, lead<strong>in</strong>g to 1836<br />

be<strong>in</strong>g real, and identical to the Hamilton-Jacobi result, equation 1829. The fact that equation 1835<br />

equals the Hamilton-Jacobi equation <strong>in</strong> the limit ~ → 0, illustrates the close analogy between the waveparticle<br />

duality of the classical Hamilton-Jacobi theory, and de Broglie’s wave-particle duality <strong>in</strong> Schröd<strong>in</strong>ger’s<br />

quantum wave-mechanics representation.<br />

The Schröd<strong>in</strong>ger approach was accepted <strong>in</strong> 1925 and exploited extensively with tremendous success, s<strong>in</strong>ce<br />

it is much easier to grasp conceptually than is the algebraic approach of Heisenberg. Initially there was much<br />

conflict between the proponents of these two contradictory approaches, but this was resolved by Schröd<strong>in</strong>ger<br />

who showed <strong>in</strong> 1926 that there is a formal mathematical identity between wave mechanics and matrix<br />

mechanics. That is, these quantal two representations of Hamiltonian mechanics are equivalent, even though<br />

they are built on either the Poisson bracket representation, or the Hamilton-Jacobi representation. Wave<br />

mechanics is based <strong>in</strong>timately on the quantization rule of the action variable. Heisenberg’s Uncerta<strong>in</strong>ty<br />

Pr<strong>in</strong>ciple is automatically satisfied by Schröd<strong>in</strong>ger’s wave mechanics s<strong>in</strong>ce the uncerta<strong>in</strong>ty pr<strong>in</strong>ciple is a<br />

feature of all wave motion, as described <strong>in</strong> chapter 3.<br />

In 1928 Dirac developed a relativistic wave equation which <strong>in</strong>cludes sp<strong>in</strong> as an <strong>in</strong>tegral part. This Dirac<br />

equation rema<strong>in</strong>s the fundamental wave equation of quantum mechanics. Unfortunately it is difficult to<br />

apply.<br />

Today the powerful and efficient Heisenberg representation is the dom<strong>in</strong>ant approach used <strong>in</strong> the field of<br />

physics, whereas chemists tend to prefer the more <strong>in</strong>tuitive Schröd<strong>in</strong>ger wave mechanics approach. In either<br />

case, the important role of Hamiltonian mechanics <strong>in</strong> quantum theory is undeniable.<br />

18.4 Lagrangian representation <strong>in</strong> quantum theory<br />

The classical notion of canonical coord<strong>in</strong>ates and momenta, has a simple quantum analog which has allowed<br />

the Hamiltonian theory of classical mechanics, that is based on canonical coord<strong>in</strong>ates, to serve as the<br />

foundation for the development of quantum mechanics. The alternative Lagrangian formulation for classical<br />

dynamics is described <strong>in</strong> terms of coord<strong>in</strong>ates and velocities, <strong>in</strong>stead of coord<strong>in</strong>ates and momenta. The Lagrangian<br />

and Hamiltonian formulations are closely related, and it may appear that the Lagrangian approach<br />

is more fundamental. The Lagrangian method allows collect<strong>in</strong>g together all the equations of motion and<br />

express<strong>in</strong>g them as stationary properties of the action <strong>in</strong>tegral, and thus it may appear desirable to base<br />

quantum mechanics on the Lagrangian theory of classical mechanics. Unfortunately, the Lagrangian equations<br />

of motion <strong>in</strong>volve partial derivatives with respect to coord<strong>in</strong>ates, and their velocities, and the mean<strong>in</strong>g<br />

ascribed to such derivatives is difficult <strong>in</strong> quantum mechanics. The close correspondence between Poisson<br />

brackets and the commutation rules leads naturally to Hamiltonian mechanics. However, Dirac showed that<br />

Lagrangian mechanics can be carried over to quantum mechanics us<strong>in</strong>g canonical transformations such that<br />

the classical Lagrangian is considered to be a function of coord<strong>in</strong>ates at time and + rather than of<br />

coord<strong>in</strong>ates and velocities.


18.5. CORRESPONDENCE PRINCIPLE 493<br />

The motivation for Feynman’s 1942 Ph.D thesis, entitled “The Pr<strong>in</strong>ciple of Least Action <strong>in</strong> Quantum<br />

<strong>Mechanics</strong>”, was to quantize the classical action at a distance <strong>in</strong> electrodynamics. This theory adopted an<br />

overall space-time viewpo<strong>in</strong>t for which the classical Hamiltonian approach, as used <strong>in</strong> conventional formulations<br />

of quantum mechanics, is <strong>in</strong>applicable. Feynman used the Lagrangian, plus the pr<strong>in</strong>ciple of least<br />

action, to underlie his development of quantum field theory. To paraphrase Feynman’s Nobel Lecture, he<br />

used a physical approach that is quite different from the customary Hamiltonian po<strong>in</strong>t of view for which the<br />

system is discussed <strong>in</strong> great detail as a function of time. That is, you have the field at this moment, then a<br />

differential equation gives you the field at a later moment and so on; that is, the Hamiltonian approach is a<br />

time differential method. In Feynman’s least-action approach the action describes the character of the path<br />

throughout all of space and time. The behavior of nature is determ<strong>in</strong>ed by say<strong>in</strong>g that the whole space-time<br />

path has a certa<strong>in</strong> character. The use of action <strong>in</strong>volves both advanced and retarded terms that make it<br />

difficult to transform back to the Hamiltonian form. The Feynman space-time approach is far beyond the<br />

scope of this course. This topic will be developed <strong>in</strong> advanced graduate courses on quantum field theory.<br />

18.5 Correspondence Pr<strong>in</strong>ciple<br />

The Correspondence Pr<strong>in</strong>ciple implies that any new theory <strong>in</strong> physics must reduce to preced<strong>in</strong>g theories<br />

that have been proven to be valid. For example, E<strong>in</strong>ste<strong>in</strong>’s Special Theory of Relativity satisfies the Correspondence<br />

Pr<strong>in</strong>ciple s<strong>in</strong>ce it reduces to classical mechanics for velocities small compared with the velocity<br />

of light. Similarly, the General Theory of Relativity reduces to Newton’s Law of Gravitation <strong>in</strong> the limit<br />

of weak gravitational fields. Bohr’s Correspondence Pr<strong>in</strong>ciple requires that the predictions of quantum mechanics<br />

must reproduce the predictions of classical physics <strong>in</strong> the limit of large quantum numbers. Bohr’s<br />

Correspondence Pr<strong>in</strong>ciple played a pivotal role <strong>in</strong> the development of the old quantum theory, from it’s<br />

<strong>in</strong>ception <strong>in</strong> 1912 until 1925 when the old quantum theory was superseded by the current matrix and wave<br />

mechanics representations of quantum mechanics.<br />

Quantum theory now is a well-established field of physics that is equally as fundamental as is classical<br />

mechanics. The Correspondence Pr<strong>in</strong>ciple now is used to project out the analogous classical-mechanics<br />

phenomena that underlie the observed properties of quantal systems. For example, this book has studied<br />

the classical-mechanics analogs of the observed behavior for typical quantal systems, such as the vibrational<br />

and rotational modes of the molecule, and the vibrational modes of the crystall<strong>in</strong>e lattice. The nucleus is the<br />

epitome of a many-body, strongly-<strong>in</strong>teract<strong>in</strong>g, quantal system. Example 1412 showed that there is a close<br />

correspondence between classical-mechanics predictions, and quantal predictions, for both the rotational and<br />

vibrational collective modes of the nucleus, as well as for the s<strong>in</strong>gle-particle motion of the nucleons <strong>in</strong> the<br />

nuclear mean field, such as the onset of Coriolis-<strong>in</strong>duced alignment. This use of the Correspondence Pr<strong>in</strong>ciple<br />

can provide considerable <strong>in</strong>sight <strong>in</strong>to the underly<strong>in</strong>g classical physics embedded <strong>in</strong> quantal systems.


494 CHAPTER 18. THE TRANSITION TO QUANTUM PHYSICS<br />

18.6 Summary<br />

The important po<strong>in</strong>t of this discussion is that variational formulations of classical mechanics provide a<br />

rational, and direct basis, for the development of quantum mechanics. It has been shown that the f<strong>in</strong>al form<br />

of quantum mechanics is closely related to the Hamiltonian formulation of classical mechanics. Quantum<br />

mechanics supersedes classical mechanics as the fundamental theory of mechanics <strong>in</strong> that classical mechanics<br />

only applies for situations where quantization is unimportant, and is the limit<strong>in</strong>g case of quantum mechanics<br />

when ~ → 0 which is <strong>in</strong> agreement with the Bohr’s Correspondence Pr<strong>in</strong>ciple. The Dirac relativistic theory<br />

of quantum mechanics is the ultimate quantal theory for the relativistic regime.<br />

This discussion has barely scratched the surface of the correspondence between classical and quantal<br />

mechanics, which goes far beyond the scope of this course. The goal of this chapter is to illustrate that<br />

classical mechanics, <strong>in</strong> particular, Hamiltonian mechanics, underlies much of what you will learn <strong>in</strong> your<br />

quantum physics courses. An <strong>in</strong>terest<strong>in</strong>g similarity between quantum mechanics and classical mechanics is<br />

that physicists usually use the more visual Schröd<strong>in</strong>ger wave representation <strong>in</strong> order to describe quantum<br />

physics to the non-expert, which is analogous to the similar use of Newtonian physics <strong>in</strong> classical mechanics.<br />

However, practic<strong>in</strong>g physicists <strong>in</strong>variably use the more abstract Heisenberg matrix mechanics to solve<br />

problems <strong>in</strong> quantum mechanics, analogous to widespread use of the variational approach <strong>in</strong> classical mechanics,<br />

because the analytical approaches are more powerful and have fundamental advantages. Quantal<br />

problems <strong>in</strong> molecular, atomic, nuclear, and subnuclear systems, usually <strong>in</strong>volve f<strong>in</strong>d<strong>in</strong>g the normal modes<br />

of a quantal system, that is, f<strong>in</strong>d<strong>in</strong>g the eigen-energies, eigen-functions, sp<strong>in</strong>, parity, and other observables<br />

for the discrete quantized levels. Solv<strong>in</strong>g the equations of motion for the modes of quantal systems is similar<br />

to solv<strong>in</strong>g the many-body coupled-oscillator problem <strong>in</strong> classical mechanics, where it was shown that<br />

use of matrix mechanics is the most powerful representation. It is ironic that the <strong>in</strong>troduction of matrix<br />

methods to classical mechanics is a by-product of the development of matrix mechanics by Heisenberg, Born<br />

and Jordan. This illustrates that classical mechanics not only played a pivotal role <strong>in</strong> the development of<br />

quantum mechanics, but it also has benefitted considerably from the development of quantum mechanics;<br />

that is, the synergistic relation between these two complementary branches of physics has been beneficial to<br />

both classical and quantum mechanics.<br />

Recommended read<strong>in</strong>g<br />

“Quantum <strong>Mechanics</strong>” by P.A.M. Dirac, Oxford Press, 1947,<br />

“Conceptual Development of Quantum <strong>Mechanics</strong>” by Max Jammer, Mc Graw Hill 1966.


Chapter 19<br />

Epilogue<br />

Stage 1<br />

Hamilton’s action pr<strong>in</strong>ciple<br />

Stage 2<br />

Hamiltonian<br />

Lagrangian<br />

d’ Alembert’s Pr<strong>in</strong>ciple<br />

Equations of motion<br />

Newtonian mechanics<br />

Stage 3<br />

Solution for motion<br />

Initial conditions<br />

Figure 19.1: Philosophical road map of the hierarchy of stages <strong>in</strong>volved <strong>in</strong> analytical mechanics. Hamilton’s<br />

Action Pr<strong>in</strong>ciple is the foundation of analytical mechanics. Stage 1 uses Hamilton’s Pr<strong>in</strong>ciple to derive the<br />

Lagranian and Hamiltonian. Stage 2 uses either the Lagrangian or Hamiltonian to derive the equations<br />

of motion for the system. Stage 3 uses these equations of motion to solve for the actual motion us<strong>in</strong>g<br />

the assumed <strong>in</strong>itial conditions. The Lagrangian approach can be derived directly based on d’Alembert’s<br />

Pr<strong>in</strong>ciple. Newtonian mechanics can be derived directly based on Newton’s Laws of Motion.<br />

This book has <strong>in</strong>troduced powerful analytical methods <strong>in</strong> physics that are based on applications of<br />

variational pr<strong>in</strong>ciples to Hamilton’s Action Pr<strong>in</strong>ciple. These methods were pioneered <strong>in</strong> classical mechanics<br />

by Leibniz, Lagrange, Euler, Hamilton, and Jacobi, dur<strong>in</strong>g the remarkable Age of Enlightenment, and reached<br />

full fruition at the start of the 20 century.<br />

The philosophical roadmap, shown above, illustrates the hierarchy of philosophical approaches available<br />

when us<strong>in</strong>g Hamilton’s Action Pr<strong>in</strong>ciple to derive the equations of motion of a system. The primary Stage1<br />

uses Hamilton’s Action functional, = R <br />

<br />

(q ˙q) to derive the Lagrangian, and Hamiltonian functionals.<br />

Stage1 provides the most fundamental and sophisticated level of understand<strong>in</strong>g and <strong>in</strong>volves specify<strong>in</strong>g<br />

all the active degrees of freedom, as well as the <strong>in</strong>teractions <strong>in</strong>volved. Stage2 uses the Lagrangian or Hamiltonian<br />

functionals, derived at Stage1, <strong>in</strong> order to derive the equations of motion for the system of <strong>in</strong>terest.<br />

Stage3 then uses the derived equations of motion to solve for the motion of the system, subject to a given<br />

set of <strong>in</strong>itial boundary conditions.<br />

Newton postulated equations of motion for nonrelativistic classical mechanics that are identical to those<br />

derived by apply<strong>in</strong>g variational pr<strong>in</strong>ciples to Hamilton’s Pr<strong>in</strong>ciple. However, Newton’s Laws of Motion are<br />

495


496 CHAPTER 19. EPILOGUE<br />

applicable only to nonrelativistic classical mechanics, and cannot exploit the advantages of us<strong>in</strong>g the more<br />

fundamental Hamilton’s Action Pr<strong>in</strong>ciple, Lagrangian, and Hamiltonian. Newtonian mechanics requires that<br />

all the active forces be <strong>in</strong>cluded <strong>in</strong> the equations of motion, and <strong>in</strong>volves deal<strong>in</strong>g with vector quantities which<br />

is more difficult than us<strong>in</strong>g the scalar functionals, action, Lagrangian, or Hamiltonian. Lagrangian mechanics<br />

based on d’Alembert’s Pr<strong>in</strong>ciple does not exploit the advantages provided by Hamilton’s Action Pr<strong>in</strong>ciple.<br />

Considerable advantages result from deriv<strong>in</strong>g the equations of motion based on Hamilton’s Pr<strong>in</strong>ciple,<br />

rather than bas<strong>in</strong>g them on the Newton’s postulated Laws of Motion. It is significantly easier to use variational<br />

pr<strong>in</strong>ciples to handle the scalar functionals, action, Lagrangian, and Hamiltonian, rather than start<strong>in</strong>g<br />

with Newton’s vector differential equations-of-motion. The three hierarchical stages of analytical mechanics<br />

facilitate accommodat<strong>in</strong>g extra degrees of freedom, symmetries, constra<strong>in</strong>ts, and other <strong>in</strong>teractions. For<br />

example, the symmetries identified by Noether’s theorem are more easily recognized dur<strong>in</strong>g the primary “action”<br />

and secondary “Hamiltonian/Lagrangian” stages, rather than at the subsequent “equations-of-motion”<br />

stage. Constra<strong>in</strong>t forces, and approximations, <strong>in</strong>troduced at the Stage1 or Stage2, are easier to implement<br />

than at the subsequent Stage3. The correspondence of Hamilton’s Action <strong>in</strong> classical and quantal mechanics,<br />

as well as relativistic <strong>in</strong>variance, are crucial advantages for us<strong>in</strong>g the analytical approach <strong>in</strong> relativistic<br />

mechanics, fluid motion, quantum, and field theory.<br />

Philosophically, Newtonian mechanics is straightforward to understand s<strong>in</strong>ce it uses vector differential<br />

equations of motion that relate the <strong>in</strong>stantaneous forces to the <strong>in</strong>stantaneous accelerations. Moreover,<br />

the concepts of momentum plus force are <strong>in</strong>tuitive to visualize, and both cause and effect are embedded<br />

<strong>in</strong> Newtonian mechanics. Unfortunately, Newtonian mechanics is <strong>in</strong>compatible with quantum physics, it<br />

violates the relativistic concepts of space-time, and fails to provide the unified description of the gravitational<br />

force plus planetary motion as geodesic motion <strong>in</strong> a four-dimensional Riemannian structure.<br />

The remarkable philosophical implications embedded <strong>in</strong> apply<strong>in</strong>g variational pr<strong>in</strong>ciples to Hamilton’s<br />

Pr<strong>in</strong>ciple, are based on the astonish<strong>in</strong>g assumption that motion of a constra<strong>in</strong>ed system <strong>in</strong> nature follows<br />

a path that m<strong>in</strong>imizes the action <strong>in</strong>tegral. As a consequence, solv<strong>in</strong>g the equations of motion is reduced<br />

to f<strong>in</strong>d<strong>in</strong>g the optimum path that m<strong>in</strong>imizes the action <strong>in</strong>tegral. The fact that nature follows optimization<br />

pr<strong>in</strong>ciples is non<strong>in</strong>tuitive, and was considered to be metaphysical by many scientists and philosophers dur<strong>in</strong>g<br />

the 19 century, which delayed full acceptance of analytical mechanics until the development of the Theory<br />

of Relativity and quantum mechanics. <strong>Variational</strong> formulations now have become the preem<strong>in</strong>ent approach<br />

<strong>in</strong> modern physics and they have toppled Newtonian mechanics from the throne of classical mechanics that<br />

it occupied for two centuries.<br />

The scope of this book extends beyond the typical classical mechanics textbook <strong>in</strong> order to illustrate<br />

how Lagrangian and Hamiltonian dynamics provides the foundation upon which modern physics is built.<br />

Knowledge of analytical mechanics is essential for the study of modern physics. The techniques and physics<br />

discussed <strong>in</strong> this book reappear <strong>in</strong> different guises <strong>in</strong> many fields, but the basic physics is unchanged illustrat<strong>in</strong>g<br />

the <strong>in</strong>tellectual beauty, the philosophical implications, and the unity of the field of physics. The breadth<br />

of physics addressed by variational pr<strong>in</strong>ciples <strong>in</strong> classical mechanics, and the underly<strong>in</strong>g unity of the field,<br />

are epitomized by the wide range of dimensions, energies, and complexity <strong>in</strong>volved. The dimensions range<br />

from as large as 10 27 to quantal analogues of classical mechanics of systems spann<strong>in</strong>g <strong>in</strong> size down to the<br />

Planck length of 162 × 10 −35 . Individual particles have been detected with k<strong>in</strong>etic energies rang<strong>in</strong>g from<br />

zero to greater than 10 15 eV. The complexity of classical mechanics spans from one body to the statistical<br />

mechanics of many-body systems. As a consequence, analytical variational methods have become the premier<br />

approach to describe systems from the very largest to the smallest, and from one-body to many-body<br />

dynamical systems.<br />

The goal of this book has been to illustrate the astonish<strong>in</strong>g power of analytical variational methods for<br />

understand<strong>in</strong>g the physics underly<strong>in</strong>g classical mechanics, as well as extensions to modern physics. However,<br />

the present narrative rema<strong>in</strong>s unf<strong>in</strong>ished <strong>in</strong> that fundamental philosophical and technical questions have<br />

not been addressed. For example, analytical mechanics is based on the validity of the assumed pr<strong>in</strong>ciple of<br />

economy. This book has not addressed the philosophical question, “is the pr<strong>in</strong>ciple of economy a fundamental<br />

law of nature, or is it a fortuitous consequence of the fundamental laws of nature?”<br />

In summary, Hamilton’s action pr<strong>in</strong>ciple, which is built <strong>in</strong>to Lagrangian and Hamiltonian mechanics,<br />

coupled with the availability of a wide arsenal of variational pr<strong>in</strong>ciples and mathematical techniques, provides<br />

a remarkably powerful approach for deriv<strong>in</strong>g the equations of motions required to determ<strong>in</strong>e the response of<br />

systems <strong>in</strong> a broad and diverse range of applications <strong>in</strong> science and eng<strong>in</strong>eer<strong>in</strong>g.


Appendix A<br />

Matrix algebra<br />

A.1 Mathematical methods for mechanics<br />

Development of classical mechanics has <strong>in</strong>volved a close and synergistic <strong>in</strong>terweav<strong>in</strong>g of physics and mathematics,<br />

that cont<strong>in</strong>ues to play a key role <strong>in</strong> these fields. The concepts of scalar and vector fields play a pivotal<br />

role <strong>in</strong> describ<strong>in</strong>g the force fields and particle motion <strong>in</strong> both the Newtonian formulation of classical mechanics<br />

and electromagnetism. Thus it is imperative that you be familiar with the sophisticated mathematical<br />

formalism used to treat multivariate scalar and vector fields <strong>in</strong> classical mechanics. Ord<strong>in</strong>ary and partial<br />

differential equations up to second order, as well as <strong>in</strong>tegration of algebraic and trigonometric functions play<br />

a major role <strong>in</strong> classical mechanics. It is assumed that you already have a work<strong>in</strong>g knowledge of differential<br />

and <strong>in</strong>tegral calculus <strong>in</strong> sufficient depth to handle this material. Computer codes, such as Mathematica,<br />

MatLab, and Maple, or symbolic calculators, can be used to obta<strong>in</strong> mathematical solutions for complicated<br />

cases.<br />

The follow<strong>in</strong>g 9 appendices provide brief summaries of matrix algebra, vector algebra, orthogonal coord<strong>in</strong>ate<br />

systems, coord<strong>in</strong>ate transformations, tensor algebra, multivariate calculus, vector differential plus<br />

<strong>in</strong>tegral calculus, Fourier analysis and time-sampled waveform analysis. The manipulation of scalar and<br />

vector fields is greatly facilitated by transform<strong>in</strong>g to orthogonal curvil<strong>in</strong>ear coord<strong>in</strong>ate systems that match<br />

the symmetries of the problem. These appendices discuss the necessity to account for the time dependence<br />

of the orthogonal unit vectors for curvil<strong>in</strong>ear coord<strong>in</strong>ate systems. It is assumed that, except for coord<strong>in</strong>ate<br />

transformations and tensor algebra, you have been <strong>in</strong>troduced to these topics <strong>in</strong> l<strong>in</strong>ear algebra and other<br />

physics courses, and thus the purpose of these appendices is to serve as a reference and brief review.<br />

A.2 Matrices<br />

Matrix algebra provides an elegant and powerful representation of multivariate operators, and coord<strong>in</strong>ate<br />

transformations that feature prom<strong>in</strong>ently <strong>in</strong> classical mechanics. For example they play a pivotal role <strong>in</strong><br />

f<strong>in</strong>d<strong>in</strong>g the eigenvalues and eigenfunctions for coupled equations that occur <strong>in</strong> rigid-body rotation, and<br />

coupled oscillator systems. An understand<strong>in</strong>g of the role of matrix mechanics <strong>in</strong> classical mechanics facilitates<br />

understand<strong>in</strong>g of the equally important role played by matrix mechanics <strong>in</strong> quantal physics.<br />

It is <strong>in</strong>terest<strong>in</strong>g that although determ<strong>in</strong>ants were used by physicists <strong>in</strong> the late 19 century, the concept<br />

of matrix algebra was developed by Arthur Cayley <strong>in</strong> England <strong>in</strong> 1855 but many of these ideas were the work<br />

of Hamilton, and the discussion of matrix algebra was buried <strong>in</strong> a more general discussion of determ<strong>in</strong>ants.<br />

Matrix algebra was an esoteric branch of mathematics, little known by the physics community, until 1925<br />

when Heisenberg proposed his <strong>in</strong>novative new quantum theory. The strik<strong>in</strong>g feature of this new theory<br />

was its representation of physical quantities by sets of time-dependent complex numbers and a peculiar<br />

multiplication rule. Max Born recognized that Heisenberg’s multiplication rule is just the standard “row<br />

times column” multiplication rule of matrix algebra; a topic that he had encountered as a young student <strong>in</strong> a<br />

mathematics course. In 1924 Richard Courant had just completed the first volume of the new text Methods<br />

of Mathematical Physics dur<strong>in</strong>g which Pascual Jordan had served as his young assistant work<strong>in</strong>g on matrix<br />

manipulation. Fortuitously, Jordan and Born happened to share a carriage on a tra<strong>in</strong> to Hanover dur<strong>in</strong>g<br />

497


498 APPENDIX A. MATRIX ALGEBRA<br />

which Jordan overheard Born talk about his problems try<strong>in</strong>g to work with matrices. Jordan <strong>in</strong>troduced<br />

himself to Born and offered to help. This led to publication, <strong>in</strong> September 1925, of the famous Born-Jordan<br />

paper[Bor25a] that gave the first rigorous formulation of matrix mechanics <strong>in</strong> physics. This was followed <strong>in</strong><br />

November by the Born-Heisenberg-Jordan sequel[Bor25b] that established a logical consistent general method<br />

for solv<strong>in</strong>g matrix mechanics problems plus a connection between the mathematics of matrix mechanics and<br />

l<strong>in</strong>ear algebra. Matrix algebra developed <strong>in</strong>to an important tool <strong>in</strong> mathematics and physics dur<strong>in</strong>g World<br />

War 2 and now it is an <strong>in</strong>tegral part of undergraduate l<strong>in</strong>ear algebra courses.<br />

Most applications of matrix algebra <strong>in</strong> this book are restricted to real, symmetric, square matrices. The<br />

size of a matrix is def<strong>in</strong>ed by the rank, which equals the row rank and column rank, i.e. the number of<br />

<strong>in</strong>dependent row vectors or column vectors <strong>in</strong> the square matrix. It is presumed that you have studied<br />

matrices <strong>in</strong> a l<strong>in</strong>ear algebra course. Thus the goal of this review is to list simple manipulation of symmetric<br />

matrices and matrix diagonalization that will be used <strong>in</strong> this course. You are referred to a l<strong>in</strong>ear algebra<br />

textbook if you need further details.<br />

Matrix def<strong>in</strong>ition<br />

A matrix is a rectangular array of numbers with rows and columns. The notation used for an element<br />

of a matrix is where designates the row and designates the column of this matrix element <strong>in</strong> the<br />

matrix A. Convention denotes a matrix A as<br />

⎛<br />

⎞<br />

11 12 1(−1) 1<br />

21 22 2(−1) 2<br />

A ≡<br />

⎜ : : : :<br />

⎟<br />

(A.1)<br />

⎝ (−1)1 (−1)2 (−1)(−1) (−1)<br />

⎠<br />

1 2 (−1) <br />

Matrices can be square, = , or rectangular 6= . Matrices hav<strong>in</strong>g only one row or column are<br />

called row or column vectors respectively, and need only a s<strong>in</strong>gle subscript label. For example,<br />

⎛ ⎞<br />

1<br />

2<br />

A =<br />

⎜ :<br />

⎟<br />

(A.2)<br />

⎝ −1<br />

⎠<br />

<br />

Matrix manipulation<br />

Matrices are def<strong>in</strong>ed to obey certa<strong>in</strong> rules for matrix manipulation as given below.<br />

1) Multiplication of a matrix by a scalar simply multiplies each matrix element by <br />

= <br />

(A.3)<br />

2) Addition of two matrices A and B hav<strong>in</strong>g the same rank, i.e. the number of columns, is given by<br />

= + <br />

(A.4)<br />

3) Multiplication of a matrix A by a matrix B is def<strong>in</strong>ed only if the number of columns <strong>in</strong> A equals the<br />

number of rows <strong>in</strong> B. The product matrix C is given by the matrix product<br />

C= A · B (A.5)<br />

=[] <br />

= X <br />

<br />

(A.6)<br />

For example, if both A and B are rank three symmetric matrices then<br />

⎛<br />

C = A · B = ⎝ ⎞ ⎛<br />

11 12 13<br />

21 22 23<br />

⎠ · ⎝ ⎞<br />

11 12 13<br />

21 22 23<br />

⎠<br />

31 32 33 31 32 33<br />

⎛<br />

= ⎝ ⎞<br />

11 11 + 12 21 + 13 31 11 12 + 12 22 + 13 32 11 13 + 12 23 + 13 33<br />

21 11 + 22 21 + 23 31 21 12 + 22 22 + 23 32 21 13 + 22 23 + 23 33<br />

⎠<br />

31 11 + 32 21 + 33 31 31 12 + 32 22 + 33 32 31 13 + 32 23 + 33 33


A.2. MATRICES 499<br />

In general, multiplication of matrices A and B is noncommutative, i.e.<br />

A · B 6= B · A<br />

In the special case when A · B = B · A then the matrices are said to commute.<br />

(A.7)<br />

Transposed matrix A <br />

The transpose of a matrix A will be denoted by A and is given by <strong>in</strong>terchang<strong>in</strong>g rows and columns, that is<br />

¡<br />

<br />

¢ = (A.8)<br />

The transpose of a column vector is a row vector. Note that older texts use the symbol à for the transpose.<br />

Identity (unity) matrix I<br />

The identity (unity) matrix I is diagonal with diagonal elements equal to 1, thatis<br />

I = <br />

(A.9)<br />

where the Kronecker delta symbol is def<strong>in</strong>ed by<br />

= 0 if 6= (A.10)<br />

= 1 if = <br />

Inverse matrix A −1<br />

If a matrix is non-s<strong>in</strong>gular, that is, its determ<strong>in</strong>ant is non-zero, then it is possible to def<strong>in</strong>e an <strong>in</strong>verse matrix<br />

A −1 . A square matrix has an <strong>in</strong>verse matrix for which the product<br />

A · A −1 = I<br />

(A.11)<br />

Orthogonal matrix<br />

Amatrixwithreal elements is orthogonal if<br />

A = A −1<br />

That is<br />

X ¡<br />

<br />

¢ = X = <br />

<br />

<br />

(A.12)<br />

(A.13)<br />

Adjo<strong>in</strong>t matrix A †<br />

For a matrix with complex elements, theadjo<strong>in</strong>t matrix, denoted by A † is def<strong>in</strong>ed as the transpose of the<br />

complex conjugate ¡<br />

A<br />

† ¢ = A∗ (A.14)<br />

Hermitian matrix<br />

The Hermitian conjugate of a complex matrix H is denoted as H † and is def<strong>in</strong>ed as<br />

H † = ¡ H ¢ ∗<br />

=(H ∗ ) <br />

(A.15)<br />

Therefore<br />

† = ∗ <br />

(A.16)<br />

AmatrixisHermitian if it is equal to its adjo<strong>in</strong>t<br />

H † = H<br />

(A.17)<br />

that is<br />

† = ∗ = <br />

(A.18)<br />

A matrix that is both Hermitian and has real elements is a symmetric matrix s<strong>in</strong>ce complex conjugation has<br />

no effect.


500 APPENDIX A. MATRIX ALGEBRA<br />

Unitary matrix<br />

Amatrixwithcomplex elements is unitary if its <strong>in</strong>verse is equal to the adjo<strong>in</strong>t matrix<br />

which is equivalent to<br />

U † = U −1<br />

U † U = I<br />

(A.19)<br />

(A.20)<br />

A unitary matrix with real elements is an orthogonal matrix as given <strong>in</strong> equation 12<br />

Trace of a square matrix A<br />

The trace of a square matrix, denoted by A, isdef<strong>in</strong>ed as the sum of the diagonal matrix elements.<br />

A =<br />

X<br />

=1<br />

<br />

(A.21)<br />

Inner product of column vectors<br />

Real vectors The generalization of the scalar (dot) product <strong>in</strong> Euclidean space is called the <strong>in</strong>ner product.<br />

Exploit<strong>in</strong>g the rules of matrix multiplication requires tak<strong>in</strong>g the transpose of the first column vector<br />

to form a row vector which then is multiplied by the second column vector us<strong>in</strong>g the conventional rules for<br />

matrix multiplication. That is, for rank vectors<br />

[X] · [Y] =<br />

⎛<br />

⎜<br />

⎝<br />

1<br />

2<br />

:<br />

<br />

⎞<br />

⎟<br />

⎠ ·<br />

⎛<br />

⎜<br />

⎝<br />

1<br />

2<br />

:<br />

<br />

⎞<br />

⎟<br />

⎠ =[X] [Y] = ¡ ¢<br />

1 2 <br />

⎛<br />

⎜<br />

⎝<br />

1<br />

2<br />

:<br />

<br />

⎞<br />

⎟<br />

<br />

⎠ = X<br />

=1<br />

<br />

(A.22)<br />

For rank =3this <strong>in</strong>ner product agrees with the conventional def<strong>in</strong>ition of the scalar product and gives a<br />

result that is a scalar. For the special case when [A] · [B] =0then the two matrices are called orthogonal.<br />

The magnitude squared of a column vector is given by the <strong>in</strong>ner product<br />

Note that this is only positive.<br />

[X] · [X] =<br />

X<br />

( ) 2 ≥ 0 (A.23)<br />

=1<br />

Complex vectors For vectors hav<strong>in</strong>g complex matrix elements the <strong>in</strong>ner product is generalized to a form<br />

that is consistent with equation 22 when the column vector matrix elements are real.<br />

[X] ∗ · [Y] =[X] † [Y] = ¡ ∗ 1 ∗ 2 ∗ −1 ∗ <br />

⎛<br />

¢<br />

⎜<br />

⎝<br />

1<br />

2<br />

:<br />

−1<br />

<br />

⎞<br />

⎟<br />

⎠ = X<br />

∗ <br />

=1<br />

(A.24)<br />

For the special case<br />

X<br />

[X] ∗ · [X] =[X] † [X] = ∗ ≥ 0<br />

=1<br />

(A.25)


A.3. DETERMINANTS 501<br />

A.3 Determ<strong>in</strong>ants<br />

Def<strong>in</strong>ition<br />

The determ<strong>in</strong>ant of a square matrix with rows equals a s<strong>in</strong>gle number derived us<strong>in</strong>g the matrix elements<br />

of the matrix. The determ<strong>in</strong>ant is denoted as det A or |A| where<br />

X<br />

|A| = ( 1 2 ) 11 22 <br />

=1<br />

(A.26)<br />

where ( 1 2 ) is the permutation <strong>in</strong>dex which is either even or odd depend<strong>in</strong>g on the number of<br />

permutations required to go from the normal order (1 2 3 ) to the sequence ( 1 2 3 ).<br />

For example for =3the determ<strong>in</strong>ant is<br />

|A| = 11 22 33 + 12 23 31 + 13 21 32 − 13 22 31 − 11 23 32 − 12 21 33<br />

(A.27)<br />

Properties<br />

1. The value of a determ<strong>in</strong>ant || =0,if<br />

(a) all elements of a row (column) are zero.<br />

(b) all elements of a row (column) are identical with, or multiples of, the correspond<strong>in</strong>g elements of<br />

another row (column).<br />

2. The value of a determ<strong>in</strong>ant is unchanged if<br />

(a) rows and columns are <strong>in</strong>terchanged.<br />

(b) a l<strong>in</strong>ear comb<strong>in</strong>ation of any number of rows is added to any one row.<br />

3. The value of a determ<strong>in</strong>ant changes sign if two rows, or any two columns, are <strong>in</strong>terchanged.<br />

4. Transpos<strong>in</strong>g a square matrix does not change its determ<strong>in</strong>ant. ¯¯A ¯¯ = |A|<br />

5. If any row (column) is multiplied by a constant factor then the value of the determ<strong>in</strong>ant is multiplied<br />

by the same factor.<br />

6. The determ<strong>in</strong>ant of a diagonal matrix equals the product of the diagonal matrix elements. That is,<br />

when = then |A| = 1 2 3 <br />

7. The determ<strong>in</strong>ant of the identity (unity) matrix |I| =1.<br />

8. The determ<strong>in</strong>ant of the null matrix, for which all matrix elements are zero, |0| =0<br />

9. A s<strong>in</strong>gular matrix has a determ<strong>in</strong>ant equal to zero.<br />

10. If each element of any row (column) appears as the sum (difference) of two or more quantities, then<br />

the determ<strong>in</strong>ant can be written as a sum (difference) of two or more determ<strong>in</strong>ants of the same order.<br />

Forexamplefororder =2<br />

¯ 11 ± 11 12 ± 12<br />

21 22<br />

¯¯¯¯ =<br />

¯ 11 12<br />

21 22<br />

¯¯¯¯ ±<br />

¯ 11 12<br />

21 22<br />

¯¯¯¯<br />

11 A determ<strong>in</strong>ant of a matrix product equals the product of the determ<strong>in</strong>ants. That is, if C = AB then<br />

|C| = |A||B|


502 APPENDIX A. MATRIX ALGEBRA<br />

Cofactor of a square matrix<br />

For a square matrix hav<strong>in</strong>g rows the cofactor is obta<strong>in</strong>ed by remov<strong>in</strong>g the row and the column<br />

and then collaps<strong>in</strong>g the rema<strong>in</strong><strong>in</strong>g matrix elements <strong>in</strong>to a square matrix with − 1 rows while preserv<strong>in</strong>g<br />

the order of the matrix elements. This is called the complementary m<strong>in</strong>or which is denoted as () . The<br />

matrix elements of the cofactor square matrix a are obta<strong>in</strong>ed by multiply<strong>in</strong>g the determ<strong>in</strong>ant of the ()<br />

complementary m<strong>in</strong>or by the phase factor (−1) + .Thatis<br />

¯¯¯ =(−1) + ()<br />

¯<br />

(A.28)<br />

The cofactor matrix has the property that<br />

X<br />

X<br />

= |A| = <br />

=1<br />

=1<br />

(A.29)<br />

Cofactors are used to expand the determ<strong>in</strong>ant of a square matrix <strong>in</strong> order to evaluate the determ<strong>in</strong>ant.<br />

Inverse of a non-s<strong>in</strong>gular matrix<br />

The ( ) matrix elements of the <strong>in</strong>verse matrix A −1 of a non-s<strong>in</strong>gular matrix A are given by the ratio of<br />

the cofactor and the determ<strong>in</strong>ant |A|, thatis<br />

−1<br />

<br />

= 1<br />

|A| <br />

Equations 28 and 29 canbeusedtoevaluatethe element of the matrix product ¡ A −1 A ¢<br />

¡<br />

A −1 A ¢ = X<br />

=1<br />

−1<br />

= 1<br />

|A|<br />

X<br />

=1<br />

= 1<br />

|A| |A| = = I <br />

(A.30)<br />

(A.31)<br />

This agrees with equation 11 that A · A −1 = I.<br />

The <strong>in</strong>verse of rank 2 or 3 matrices is required frequently when determ<strong>in</strong><strong>in</strong>g the eigen-solutions for rigidbody<br />

rotation, or coupled oscillator, problems <strong>in</strong> classical mechanics as described <strong>in</strong> chapters 11 and 12.<br />

Therefore it is convenient to list explicitly the <strong>in</strong>verse matrices for both rank 2 and rank 3 matrices.<br />

Inverse for rank 2 matrices:<br />

∙ ¸−1<br />

A −1 <br />

=<br />

= 1 ∙ ¸<br />

∙ − 1 −<br />

=<br />

|A| − ( − ) − <br />

where the determ<strong>in</strong>ant of A is written explicitly <strong>in</strong> equation 32.<br />

Inverse for rank 3 matrices:<br />

¸<br />

(A.32)<br />

A −1 =<br />

=<br />

⎡<br />

⎣ ⎤−1<br />

⎦<br />

<br />

⎡<br />

⎣<br />

1<br />

+ + <br />

⎡<br />

= 1 ⎣ <br />

<br />

|A|<br />

<br />

⎤<br />

⎦<br />

<br />

⎡<br />

= 1 ⎣ <br />

<br />

|A|<br />

<br />

=( − ) = − ( − ) =( − )<br />

= − ( − ) =( − ) = − ( − )<br />

=( − ) = − ( − ) =( − )<br />

⎤<br />

⎦<br />

⎤<br />

⎦<br />

(A.33)<br />

where the functions are equal to rank 2 determ<strong>in</strong>ants listed <strong>in</strong> equation 33.


A.4. REDUCTION OF A MATRIX TO DIAGONAL FORM 503<br />

A.4 Reduction of a matrix to diagonal form<br />

Solv<strong>in</strong>g coupled l<strong>in</strong>ear equations can be reduced to diagonalization of a matrix. Consider the matrix A<br />

operat<strong>in</strong>g on the vector X to produce a vector Y, that are expressed as components with respect to the<br />

unprimed coord<strong>in</strong>ate frame, i.e.<br />

A · X = Y<br />

(A.34)<br />

Consider that the unitary real matrix R with rank , rotates the -dimensional un-primed coord<strong>in</strong>ate<br />

frame <strong>in</strong>to the primed coord<strong>in</strong>ate frame such that A , X and Y are transformed to A 0 , X 0 and Y 0 <strong>in</strong> the<br />

rotated primed coord<strong>in</strong>ate frame. Then<br />

With respect to the primed coord<strong>in</strong>ate frame equation (34) becomes<br />

X 0 = R · X<br />

Y 0 = R · Y (A.35)<br />

R· (A · X) = R · Y (A.36)<br />

R · A · R −1 · R · X = R · Y (A.37)<br />

R · A · R −1 · X 0 = A 0 · X 0 = Y 0 (A.38)<br />

us<strong>in</strong>g the fact that the identity matrix I = R · R −1 = R · R s<strong>in</strong>ce the rotation matrix <strong>in</strong> dimensions is<br />

orthogonal.<br />

Thus we have that the rotated matrix<br />

A 0 = R · A · R <br />

(A.39)<br />

Let us assume that this transformed matrix is diagonal, then it can be written as the product of the unit<br />

matrix I and a vector of scalar numbers called the characteristic roots as<br />

A 0 = R · A · R = I<br />

(A.40)<br />

us<strong>in</strong>g the fact that R = R −1 then gives<br />

R · (I) =A 0·R <br />

(A.41)<br />

Let both sides of equation 41 act on X 0 which gives<br />

I·X 0 = A 0·X 0<br />

(A.42)<br />

or £<br />

I−A<br />

0 ¤ X 0 = 0 (A.43)<br />

This represents a set of homogeneous l<strong>in</strong>ear algebraic equations <strong>in</strong> unknowns X 0 where is a set of<br />

characteristic roots, (eigenvalues) with correspond<strong>in</strong>g eigenfunctions X 0 Ignor<strong>in</strong>g the trivial case of X 0 be<strong>in</strong>g<br />

zero, then (43) requires that the secular determ<strong>in</strong>ant of the bracket be zero, that is<br />

¯<br />

¯I−A<br />

0¯¯ = 0<br />

(A.44)<br />

The determ<strong>in</strong>ant can be expanded and factored <strong>in</strong>to the form<br />

( − 1 )( − 2 )( − 3 ) ( − )=0 (A.45)<br />

where the eigenvalues are = 1 2 of the matrix A 0 <br />

The eigenvectors X 0 correspond<strong>in</strong>g to each eigenvalue are determ<strong>in</strong>ed by substitut<strong>in</strong>g a given eigenvalue<br />

<strong>in</strong>to the relation<br />

X 0 · A 0·X 0 =[ ]<br />

(A.46)<br />

If all the eigenvalues are dist<strong>in</strong>ct, i.e. different, then this set of equations completely determ<strong>in</strong>es the ratio<br />

of the components of each eigenvector along the axes of the coord<strong>in</strong>ate frame. However, when two or more


504 APPENDIX A. MATRIX ALGEBRA<br />

eigenvalues are identical, then the reduction to a true diagonal form is not possible and one has the freedom<br />

to select an appropriate eigenvector that is orthogonal to the rema<strong>in</strong><strong>in</strong>g axes.<br />

In summary, the matrix can only be fully diagonalized if (a) all the eigenvalues are dist<strong>in</strong>ct, (b) the real<br />

matrix is symmetric, (c) it is unitary.<br />

A frequent application of matrices <strong>in</strong> classical mechanics is for solv<strong>in</strong>g a system of homogeneous l<strong>in</strong>ear<br />

equations of the form<br />

11 1 + 12 2 + 1 = 0<br />

11 1 + 12 2 + 1 = 0<br />

(A.47)<br />

= <br />

1 1 + 2 2 + = 0<br />

Mak<strong>in</strong>g the follow<strong>in</strong>g def<strong>in</strong>itions<br />

⎛<br />

A = ⎜<br />

⎝<br />

⎞<br />

11 12 1<br />

21 22 2<br />

⎟<br />

⎠<br />

1 2 <br />

⎛<br />

X = ⎜<br />

⎝<br />

Then the set of l<strong>in</strong>ear equations can be written <strong>in</strong> a compact form us<strong>in</strong>g the matrices<br />

1<br />

2<br />

<br />

<br />

⎞<br />

⎟<br />

⎠<br />

(A.48)<br />

(A.49)<br />

A · X =0<br />

(A.50)<br />

which can be solved us<strong>in</strong>g equation (43). Ensure that you are able to diagonalize a matrices with rank<br />

2 and 3. You can use Mathematica, Maple, MatLab, or other such mathematical computer programs to<br />

diagonalize larger matrices.<br />

A.1 Example: Eigenvalues and eigenvectors of a real symmetric matrix<br />

Consider the matrix<br />

The secular determ<strong>in</strong>ant is given by (42)<br />

¯<br />

⎛<br />

A = ⎝ 0 1 0 ⎞<br />

1 0 0 ⎠<br />

0 0 0<br />

− 1 0<br />

1 − 0<br />

0 0 −<br />

¯ =0<br />

This expands to<br />

−( +1)( − 1) = 0<br />

Thus the three eigen values are = −1 0 1<br />

To f<strong>in</strong>d each eigenvectors we substitute the correspond<strong>in</strong>g eigenvalue <strong>in</strong>to equation (48) <br />

⎛<br />

⎝ − 1 0 ⎞ ⎛<br />

1 − 0 ⎠ ⎝ ⎞ ⎛<br />

⎠ = ⎝ 0 ⎞<br />

0 ⎠<br />

0 0 − 0<br />

The eigenvalue = −1 yields + =0and =0 Thus the eigen vector is 1 =( √ 1<br />

2<br />

√ −1<br />

2<br />

0). The<br />

eigenvalue =0yields =0and =0 Thus the eigen vector is 2 =(0 0 1). The eigenvalue =1<br />

yields − + =0and =0 Thus the eigen vector is 3 =( √ 1 1<br />

2<br />

√2 0). The orthogonality of these three<br />

eigen vectors, which correspond to three dist<strong>in</strong>ct eigenvalues, can be verified.


Appendix B<br />

Vector algebra<br />

B.1 L<strong>in</strong>ear operations<br />

The important force fields <strong>in</strong> classical mechanics, namely, gravitation, electric, and magnetic, are vector<br />

fields that have a position-dependent magnitude and direction. Thus, it is useful to summarize the algebra<br />

of vector fields.<br />

Avectora has both a magnitude || andadirectiondef<strong>in</strong>ed by the unit vector ê , that is, the vector<br />

can be written as a bold character a where<br />

a = · ê <br />

(B.1)<br />

where by convention the implied modulus sign is omitted. The hat symbol on the vector ê designates that<br />

this is a unit vector with modulus |ê | =1.<br />

Vector force fields are assumed to be l<strong>in</strong>ear, and consequently they obey the pr<strong>in</strong>ciple of superposition,<br />

are commutative, associative, and distributive as illustrated below for three vectors a b c plus a scalar<br />

multiplier <br />

a ± b = ±b + a (B.2)<br />

a+(b + c) = (a + b)+c<br />

(a + b) = a+b<br />

The manipulation of vectors is greatly facilitated by use of components along an orthogonal coord<strong>in</strong>ate<br />

system def<strong>in</strong>ed by three orthogonal unit vectors (ê 1 ê 2 ê 3 ) . For example the cartesian coord<strong>in</strong>ate system<br />

is def<strong>in</strong>ed by three unit vectors which, by convention, are called (î ĵ ˆk).<br />

B.2 Scalar product<br />

Multiplication of two vectors can produce a 9−component tensor that can be represented by a 3 × 3 matrix<br />

as discussed <strong>in</strong> appendix . There are two special cases for vector multiplication that are important for<br />

vector algebra; the first is the scalar product, and the second is the vector product.<br />

The scalar product of two vectors is def<strong>in</strong>ed to be<br />

a · b = |||| cos <br />

(B.3)<br />

where is the angle between the two vectors. It is a scalar and thus is <strong>in</strong>dependent of the orientation of<br />

the coord<strong>in</strong>ate axis system. Note that the scalar product commutes, is distributive, and associative with a<br />

scalar multiplier, that is<br />

a · b = b · a (B.4)<br />

a· (b + c) = a · b + a · c<br />

(a) ·b = (b · a)<br />

Note that a · a = || 2 and if a and b are perpendicular then cos =0and thus a · b =0<br />

505


506 APPENDIX B. VECTOR ALGEBRA<br />

If the three unit vectors (ê 1 ê 2 ê 3 ) form an orthonormal basis, that is, they are orthogonal unit vectors,<br />

then from equations 3 and 4<br />

ê · ê = <br />

(B.5)<br />

If â is the unit vector for the vector a then the scalar product of a vector a with one of these unit vectors<br />

ê gives the cos<strong>in</strong>e of the angle between the vector a and ê ,thatis<br />

a · ê 1 = || (â · ê 1 )=|| cos (B.6)<br />

a · ê 2 = || (â · ê 2 )=|| cos <br />

a · ê 3 = || (â · ê 3 )=|| cos <br />

where the cos<strong>in</strong>es are called the direction cos<strong>in</strong>es s<strong>in</strong>ce they def<strong>in</strong>e the direction of the vector a with respect<br />

to each orthogonal basis unit vector. Moreover, a · ê 1 = || â · ê 1 = || cos is the component of a along the<br />

ê 1 axis. Thus the three components of the vector a is fully def<strong>in</strong>ed by the magnitude || and the direction<br />

cos<strong>in</strong>es, correspond<strong>in</strong>g to the angles . Thatis,<br />

1 = || (â · ê 1 )=|| cos (B.7)<br />

2 = || (â · ê 2 )=|| cos <br />

3 = || (â · ê 3 )=|| cos <br />

If the three unit vectors (ê 1 ê 2 ê 3 ) form an orthonormal basis then the vector is fully def<strong>in</strong>ed by<br />

a = 1 ê 1 + 2 ê 2 + 3 ê 3<br />

(B.8)<br />

Consider two vectors<br />

a = 1 ê 1 + 2 ê 2 + 3 ê 3<br />

b = 1 ê 1 + 2 ê 2 + 3 ê 3<br />

Then us<strong>in</strong>g 5<br />

a · b = 1 1 + 2 2 + 3 3 = |||| cos <br />

(B.9)<br />

where is the angle between the two vectors. In particular, s<strong>in</strong>ce the direction cos<strong>in</strong>e cos = 1<br />

||<br />

,then<br />

equation 9 gives<br />

cos =cos cos +cos cos +cos cos <br />

(B.10)<br />

Note that when =0then 10 gives<br />

cos 2 +cos 2 +cos 2 =1<br />

(B.11)<br />

B.3 Vector product<br />

The vector product of two vectors is def<strong>in</strong>ed to be<br />

c = a × b = |||| s<strong>in</strong> ˆn<br />

(B.12)<br />

where is the angle between the³<br />

vectors´<br />

and ˆn is a unit vector perpendicular to the plane def<strong>in</strong>ed by a<br />

and b such that the unit vectors â ˆb ˆn obey a right-handed screw rule. The vector product acts like a<br />

pseudovector which comprises a normal vector multiplied by a sign factor that depends on the handedness<br />

of the system as described <strong>in</strong> appendix 3.<br />

The components of c are def<strong>in</strong>ed by the relation<br />

≡ X <br />

<br />

(B.13)<br />

where the (Levi-Civita) permutation symbol has the follow<strong>in</strong>g properties<br />

=0 if an <strong>in</strong>dex is equal to any another <strong>in</strong>dex<br />

=+1 if form an even permutation of 1 2 3<br />

= −1 if form an odd permutation of 1 2 3<br />

(B.14)


B.4. TRIPLE PRODUCTS 507<br />

For example, if the three unit vectors (ê 1 ê 2 ê 3 ) form an orthonormal basis, then ê ≡ P ê ê , i.e.<br />

ê 1 × ê 2 = ê 3 ê 2 × ê 3 = ê 1 ê 3 × ê 1 = ê 2 (B.15)<br />

ê 2 × ê 1 = −ê 3 ê 3 × ê 2 = −ê 1 ê 1 × ê 3 = −ê 2 (B.16)<br />

ê 1 × ê 1 = 0 ê 2 × ê 2 = 0 ê 3 × ê 0 = 0 (B.17)<br />

The vector product anticommutes <strong>in</strong> that<br />

a × b = −b × a<br />

(B.18)<br />

However, it is distributive and associative with a scalar multiplier<br />

Note that when s<strong>in</strong> =0then a × b =0and <strong>in</strong> particular, a × a =0<br />

Consider two vectors<br />

a× (b + c) = a × b + a × c (B.19)<br />

(a) ×b = (a × b) (B.20)<br />

a = 1 ê 1 + 2 ê 2 + 3 ê 3<br />

b = 1 ê 1 + 2 ê 2 + 3 ê 3<br />

Then us<strong>in</strong>g equations 12 and 15 − 17<br />

ê 1 ê 2 ê 3<br />

a × b= |||| s<strong>in</strong> =<br />

1 2 3 = ê 1 ( 2 3 − 3 2 )+ê 2 ( 3 1 − 1 3 )+ê 3 ( 1 2 − 2 1 )<br />

¯ 1 2 3<br />

¯¯¯¯¯¯<br />

where is the angle between the two vectors and the determ<strong>in</strong>ant is evaluated for the top row. Examples of<br />

vector products are torque N = r × F, angular momentum L = r × p, and the magnetic force F = v × B.<br />

B.4 Triple products<br />

The follow<strong>in</strong>g scalar and vector triple products can be formed from the product of three vectors and are<br />

used frequently.<br />

Scalar triple products<br />

There are several permutations of scalar triple products of three vectors [a b c] that are identical.<br />

a· (b × c) =c· (a × b) =b· (c × a) =(a × b) · c = −a· (c × b) (B.21)<br />

That is, the scalar product is <strong>in</strong>variant to cyclic permutations of the three vectors but changes sign for<br />

<strong>in</strong>terchange of two vectors. The scalar product is unchanged by swapp<strong>in</strong>g the scalar ()and vector ().<br />

Because of the symmetry the scalar triple product can be denoted as [a b c] and<br />

[a b c] 0 if [a b c] is right-handed<br />

[a b c] = 0 if [a b c] is coplanar (B.22)<br />

[a b c] 0 if [a b c] is left-handed<br />

The scalar triple product can be written <strong>in</strong> terms of the components us<strong>in</strong>g a determ<strong>in</strong>ant<br />

1 2 3<br />

[a b c] =<br />

1 2 3<br />

¯ 1 2 3<br />

¯¯¯¯¯¯<br />

(B.23)


508 APPENDIX B. VECTOR ALGEBRA<br />

Vector triple product<br />

The vector triple product a× (b × c) is a vector. S<strong>in</strong>ce (b × c) is perpendicular to the plane of b c, then<br />

a× (b × c) must lie <strong>in</strong> the plane conta<strong>in</strong><strong>in</strong>g b c. Therefore the triple product can be expanded <strong>in</strong> terms of<br />

b c, as given by the follow<strong>in</strong>g identity<br />

a × (b × c) =(a · c) b − (a · b) c<br />

(B.24)<br />

Problems<br />

1. Partition the follow<strong>in</strong>g exercises among the group. Once you have completed your problem, check with a<br />

classmate before writ<strong>in</strong>g it on the board. After you have verified that you have found the correct solution,<br />

write your answer <strong>in</strong> the space provided on the board, tak<strong>in</strong>g care to <strong>in</strong>clude the steps that you used to arrive<br />

at your solution. The follow<strong>in</strong>g <strong>in</strong>formation is needed.<br />

a ⎛=3i +2j − 9k b = −2i +3k c = −2i + j − 6k d = i +9j +4k<br />

E = ⎝ 2 7 −4 ⎞<br />

⎛<br />

µ <br />

3 1 −2 ⎠ 3 4<br />

F = G = ⎝ 2 −4 ⎞ ⎛<br />

⎞<br />

−8 −1 −3<br />

7 1 ⎠ H = ⎝ −4 2 −2 ⎠<br />

5 6<br />

−2 0 5<br />

−1 1<br />

−1 0 0<br />

Calculate each of the follow<strong>in</strong>g<br />

1 |a − (b + 3c)| 7 (EH) <br />

2 Component of c along a 8 |HE|<br />

3 Angle between c and d 9 EHG<br />

4 (b × d) · a 10 EG − HG<br />

5 (b × d) × a 11 EH − H E <br />

6 b× (d × a) 12 F −1<br />

2. For what values of are the vectors A =2ˆ − 2ˆ + ˆ and B = ˆ +2ˆ +2ˆ perpendicular?<br />

3. Show that the triple scalar product ( × ) · can be written as<br />

1 2 3<br />

(A × B) · C =<br />

1 2 3<br />

¯ 1 2 3<br />

¯¯¯¯¯¯<br />

Show also that the product is unaffected by <strong>in</strong>terchange of the scalar and vector product operations or by<br />

change <strong>in</strong> the order of as long as they are <strong>in</strong> cyclic order, that is<br />

(A × B) · C = A · (B × C) =B · (C × A) =(C × A) · B<br />

Therefore we may use the notation to denote the triple scalar product. F<strong>in</strong>ally give a geometric <strong>in</strong>terpretation<br />

of by comput<strong>in</strong>g the volume of the parallelepiped def<strong>in</strong>ed by the three vectors A B C


Appendix C<br />

Orthogonal coord<strong>in</strong>ate systems<br />

The methods of vector analysis provide a convenient representation of physical laws. However, the manipulation<br />

of scalar and vector fields is greatly facilitated by use of components with respect to an orthogonal<br />

coord<strong>in</strong>ate system.<br />

C.1 Cartesian coord<strong>in</strong>ates ( )<br />

Cartesian coord<strong>in</strong>ates (rectangular) provide the simplest orthogonal rectangular coord<strong>in</strong>ate system. The<br />

unit vectors specify<strong>in</strong>g the direction along the three orthogonal axes are taken to be (î ĵ ˆk). In cartesian<br />

coord<strong>in</strong>ates scalar and vector functions are written as<br />

= ( ) (C.1)<br />

r = î+ĵ+ˆk (C.2)<br />

Calculation of the time derivatives of the position vector is especially simple us<strong>in</strong>g cartesian coord<strong>in</strong>ates<br />

because the unit vectors (î ĵ ˆk) are constant and <strong>in</strong>dependent <strong>in</strong> time. That is;<br />

î<br />

= ĵ<br />

= ˆk<br />

=0<br />

S<strong>in</strong>ce the time derivatives of the unit vectors are all zero then the velocity ṙ = r<br />

<br />

derivatives of and . Thatis,<br />

ṙ = ˙î+ ˙ĵ+ ˙ˆk<br />

(C.3)<br />

Similarly the acceleration is given by<br />

¨r =¨î+¨ĵ+¨ˆk<br />

reduces to the partial time<br />

(C.4)<br />

C.2 Curvil<strong>in</strong>ear coord<strong>in</strong>ate systems<br />

There are many examples <strong>in</strong> physics where the symmetry of the problem makes it more convenient to solve<br />

motion at a po<strong>in</strong>t ( ) us<strong>in</strong>g non-cartesian curvil<strong>in</strong>ear coord<strong>in</strong>ate systems. For example, problems<br />

hav<strong>in</strong>g spherical symmetry are most conveniently handled us<strong>in</strong>g a spherical coord<strong>in</strong>ate system ( )<br />

with the orig<strong>in</strong> at the center of spherical symmetry. Such problems occur frequently <strong>in</strong> electrostatics and<br />

gravitation; e.g. solutions of the atom, or planetary systems. Note that a cartesian coord<strong>in</strong>ate system still<br />

is required to def<strong>in</strong>e the orig<strong>in</strong> plus the polar and azimuthal angles Us<strong>in</strong>g spherical coord<strong>in</strong>ates for<br />

a spherically symmetry system allows the problem to be factored <strong>in</strong>to a cyclic angular part, the solution<br />

which <strong>in</strong>volves spherical harmonics that are common to all such spherically-symmetric problems, plus a<br />

one-dimensional radial part that conta<strong>in</strong>s the specifics of the particular spherically-symmetric potential.<br />

Similarly, for problems <strong>in</strong>volv<strong>in</strong>g cyl<strong>in</strong>drical symmetry, it is much more convenient to use a cyl<strong>in</strong>drical<br />

coord<strong>in</strong>ate system ( ). Aga<strong>in</strong> it is necessary to use a cartesian coord<strong>in</strong>ate system to def<strong>in</strong>e the orig<strong>in</strong><br />

and angle . Motion <strong>in</strong> a plane can be handled us<strong>in</strong>g two dimensional polar coord<strong>in</strong>ates.<br />

509


510 APPENDIX C. ORTHOGONAL COORDINATE SYSTEMS<br />

Curvil<strong>in</strong>ear coord<strong>in</strong>ate systems <strong>in</strong>troduce a complication <strong>in</strong> that the unit vectors are time dependent <strong>in</strong><br />

contrast to cartesian coord<strong>in</strong>ate system where the unit vectors (î ĵ ˆk) are <strong>in</strong>dependent and constant <strong>in</strong> time.<br />

The <strong>in</strong>troduction of this time dependence warrants further discussion.<br />

Each of the three axes <strong>in</strong> curvil<strong>in</strong>ear coord<strong>in</strong>ate systems can be expressed <strong>in</strong> cartesian coord<strong>in</strong>ates<br />

( ) as surfaces of constant given by the function<br />

= ( )<br />

(C.5)<br />

where =1 2 or 3. An element of length perpendicular to the surface is the distance between the<br />

surfaces and + whichcanbeexpressedas<br />

= <br />

(C.6)<br />

where is a function of ( 1 2 3 ). In cartesian coord<strong>in</strong>ates 1 , 2 and 3 are all unity. The unit-length<br />

vectors ˆ 1 , ˆ 2 , ˆ 3 , are perpendicular to the respective 1 2 3 surfaces, and are oriented to have <strong>in</strong>creas<strong>in</strong>g<br />

<strong>in</strong>dices such that ˆq 1 ׈q 2 = ˆq 3 . The correspondence of the curvil<strong>in</strong>ear coord<strong>in</strong>ates, unit vectors, and transform<br />

coefficients to cartesian, polar, cyl<strong>in</strong>drical and spherical coord<strong>in</strong>ates is given <strong>in</strong> table 1<br />

Curvil<strong>in</strong>ear 1 2 3 ˆq 1 ˆq 2 ˆq 3 1 2 3<br />

Cartesian ˆ ˆ ˆk 1 1 1<br />

Polar ˆr ˆθ 1 <br />

Cyl<strong>in</strong>drical ˆρ ˆϕ ẑ 1 1<br />

Spherical ˆr ˆθ ˆϕ 1 <br />

Table 1: Curvil<strong>in</strong>ear coord<strong>in</strong>ates<br />

The differential distance and volume elements are given by<br />

s = 1ˆq 1 + 2ˆq 2 + 3ˆq 3 = 1 1ˆq 1 + 2 2ˆq 2 + 3 3ˆq 3 (C.7)<br />

= 1 2 3 = 1 2 3 ( 1 2 3 ) (C.8)<br />

These are evaluated below for polar, cyl<strong>in</strong>drical, and spherical coord<strong>in</strong>ates.<br />

C.2.1 Two-dimensional polar coord<strong>in</strong>ates ( )<br />

The complication and implications of time-dependent unit vectors are best illustrated by consider<strong>in</strong>g twodimensional<br />

polar coord<strong>in</strong>ates which is the simplest curvil<strong>in</strong>ear coord<strong>in</strong>ate system. Polar coord<strong>in</strong>ates are a<br />

special case of cyl<strong>in</strong>drical coord<strong>in</strong>ates, when is held fixed, or a special case of spherical coord<strong>in</strong>ate system,<br />

when is held fixed.<br />

Consider the motion of a po<strong>in</strong>t as it moves along a curve s() such that <strong>in</strong> the time <strong>in</strong>terval it moves<br />

from (1) to (2) as shown <strong>in</strong> figure 2. The two-dimensional polar coord<strong>in</strong>ates have unit vectors ˆr, ˆθ,<br />

which are orthogonal and change from ˆr 1 , ˆθ 1 ,toˆr 2 , ˆθ 2 ,<strong>in</strong>thetime Notethatforthesepolarcoord<strong>in</strong>ates<br />

the angle unit vector ˆθ is taken to be tangential to the rotation s<strong>in</strong>ce this is the direction of motion of a<br />

po<strong>in</strong>t on the circumference at radius .<br />

The net changes shown <strong>in</strong> figure of table 2 are<br />

ˆr = ˆr 2 − ˆr 1 = ˆr = |ˆr| ˆθ =ˆθ<br />

(C.9)<br />

s<strong>in</strong>ce the unit vector ˆr is a constant with |ˆr| =1.Notethatthe<strong>in</strong>f<strong>in</strong>itessimal ˆr is perpendicular to the unit<br />

vector ˆr, thatis,ˆr po<strong>in</strong>ts <strong>in</strong> the tangential direction ˆθ<br />

Similarly, the <strong>in</strong>f<strong>in</strong>itessimal<br />

ˆθ = ˆθ 2 − ˆθ 1 = ˆθ = −ˆr<br />

(C.10)<br />

which is perpendicular to the tangential ˆθ unit vector and therefore po<strong>in</strong>ts <strong>in</strong> the direction −ˆr . The m<strong>in</strong>us<br />

sign causes −ˆr to be directed <strong>in</strong> the opposite direction to ˆr.


C.2. CURVILINEAR COORDINATE SYSTEMS 511<br />

The net distance element s is given by<br />

s =ˆr + dˆr =ˆr + ˆθ<br />

(C.11)<br />

This agrees with the prediction obta<strong>in</strong>ed us<strong>in</strong>g table 1<br />

Thetimederivativesoftheunitvectorsaregivenbyequations(9) and (10) to be,<br />

ˆr<br />

<br />

ˆθ<br />

<br />

= <br />

ˆθ<br />

= − <br />

ˆr<br />

(C.12)<br />

(C.13)<br />

Note that the time derivatives of unit vectors are perpendicular to the correspond<strong>in</strong>g unit vector, and the<br />

unit vectors are coupled.<br />

Consider that the velocity v is expressed as<br />

v= r<br />

= ˆr<br />

(ˆr) = + <br />

ˆr<br />

<br />

= ˙ˆr + ˙ˆθ<br />

(C.14)<br />

The velocity is resolved <strong>in</strong>to a radial component ˙ and an angular, transverse, component ˙.<br />

Similarly the acceleration is given by<br />

a = v<br />

= ˆr+ ˙ ˆr<br />

˙<br />

+ <br />

<br />

2´<br />

=<br />

³¨ − ˙ ˆr +<br />

˙ ˆθ<br />

˙ˆθ+ ˆθ+ ˙<br />

<br />

³<br />

¨ +2˙ ˙´<br />

ˆθ (C.15)<br />

where the ˙ 2ˆr term is the effective centripetal acceleration while the 2 ˙ ˙ˆθ term is called the Coriolis term.<br />

For the case when ˙ =¨ =0, then the first bracket <strong>in</strong> 15 is the centripetal acceleration while the second<br />

bracket is the tangential acceleration.<br />

This discussion has shown that <strong>in</strong> contrast to the time <strong>in</strong>dependence of the cartesian unit basis vectors,<br />

the unit basis vectors for curvil<strong>in</strong>ear coord<strong>in</strong>ates are time dependent which leads to components of the velocity<br />

and acceleration <strong>in</strong>volv<strong>in</strong>g coupled coord<strong>in</strong>ates.<br />

Coord<strong>in</strong>ates<br />

Distance element<br />

Area element<br />

Unit vectors<br />

Time derivatives<br />

of unit vectors<br />

Velocity<br />

K<strong>in</strong>etic energy<br />

Acceleration a =<br />

<br />

s = ˆr + ˆθ<br />

= <br />

ˆr = ˆ cos + ˆ s<strong>in</strong> <br />

ˆθ = −ˆ s<strong>in</strong> + ˆ cos <br />

ˆr<br />

= ˙ˆθ<br />

ˆ<br />

= −˙ˆr<br />

v =<br />

³<br />

˙ˆr + ˙ˆθ<br />

<br />

2<br />

˙ 2 2<br />

2´<br />

+ ˙<br />

2´ ³¨ − ˙ ˆr<br />

³<br />

˙´<br />

+ ¨ +2˙ ˆθ<br />

Table 2: Differential relations plus a diagram of the unit vectors for 2-dimensional polar coord<strong>in</strong>ates.


512 APPENDIX C. ORTHOGONAL COORDINATE SYSTEMS<br />

C.2.2 Cyl<strong>in</strong>drical Coord<strong>in</strong>ates ( )<br />

The three-dimensional cyl<strong>in</strong>drical coord<strong>in</strong>ates ( ) are obta<strong>in</strong>ed by add<strong>in</strong>g the motion along the symmetry<br />

axis ẑ to the case for polar coord<strong>in</strong>ates. The unit basis vectors are shown <strong>in</strong> Table 3 where the angular<br />

unit vector ˆφ is taken to be tangential correspond<strong>in</strong>g to the direction a po<strong>in</strong>t on the circumference would<br />

move. The distance and volume elements, the cartesian coord<strong>in</strong>ate components of the cyl<strong>in</strong>drical unit<br />

basis vectors, and the unit vector time derivatives are shown <strong>in</strong> Table 3. The time dependence of the<br />

unit vectors is used to derive the acceleration. As for the two-dimensional polar coord<strong>in</strong>ates, the ˆρ and ˆθ<br />

direction components of the acceleration for cyl<strong>in</strong>drical coord<strong>in</strong>ates are coupled functions of ˙ ¨ ˙ and ¨.<br />

Coord<strong>in</strong>ates<br />

Distance element<br />

Volume element<br />

Unit vectors<br />

Time derivatives<br />

of unit vectors<br />

Velocity<br />

K<strong>in</strong>etic energy<br />

<br />

s = ˆρ + ˆφ + ẑ<br />

= <br />

ˆρ = ˆ cos + ˆ s<strong>in</strong> <br />

ˆφ = −ˆ s<strong>in</strong> + ˆ cos <br />

ẑ = ˆk<br />

ˆ<br />

= ˙ˆφ<br />

ˆ<br />

= − ˙ˆρ<br />

ẑ<br />

=0<br />

v = ˙ˆρ + ˙ˆφ + ˙ẑ<br />

³˙ 2 + 2 ˙ 2´ 2 + ˙<br />

<br />

2<br />

Acceleration a =<br />

+<br />

2´ ³¨ − ˙ ˆρ<br />

³<br />

¨ +2˙ ˙´<br />

ˆφ +¨ẑ<br />

Table 3: Differential relations plus a diagram of the unit vectors for cyl<strong>in</strong>drical coord<strong>in</strong>ates.<br />

C.2.3 Spherical Coord<strong>in</strong>ates ( )<br />

The three dimensional spherical coord<strong>in</strong>ates, can be treated the same way as for cyl<strong>in</strong>drical coord<strong>in</strong>ates. The<br />

unit basis vectors are shown <strong>in</strong> Table 4 where the angular unit vectors ˆθ and ˆφ are taken to be tangential<br />

correspond<strong>in</strong>g to the direction a po<strong>in</strong>t on the circumference moves for a positive rotation angle.<br />

Coord<strong>in</strong>ates<br />

<br />

Distance element = ˆr + ˆθ + s<strong>in</strong> ˆφ<br />

Volume element<br />

= 2 s<strong>in</strong> <br />

Unit vectors ˆr = ˆ s<strong>in</strong> cos + ˆ s<strong>in</strong> s<strong>in</strong> + ˆk cos <br />

ˆθ = ˆ cos cos + ˆ cos s<strong>in</strong> − ˆk s<strong>in</strong> <br />

ˆφ = −ˆ s<strong>in</strong> + ˆ cos <br />

ˆr<br />

Time derivatives<br />

= ˆθ ˙ + ˆφ ˙ s<strong>in</strong> <br />

ˆ<br />

of unit vectors<br />

= −ˆr˙ + ˆφ ˙ cos <br />

ˆ<br />

= −ˆr ˙ s<strong>in</strong> − ˆθ ˙ cos <br />

Velocity<br />

v =<br />

³<br />

˙ˆr + ˙ˆθ + ˙ s<strong>in</strong> ˆφ<br />

<br />

K<strong>in</strong>etic energy<br />

2<br />

˙ 2 + 2 ˙ 2´<br />

2<br />

+ 2 s<strong>in</strong> 2 ˙<br />

³<br />

Acceleration a = ¨ − ˙ 2 − ˙<br />

´<br />

2<br />

s<strong>in</strong> 2 ˆr<br />

³<br />

+ ¨ +2˙ ˙ − ˙ 2 s<strong>in</strong> cos ´<br />

ˆθ<br />

³<br />

+ ¨ s<strong>in</strong> +2˙ ˙ s<strong>in</strong> +2 ˙ ˙ cos ´<br />

ˆφ<br />

Table 4 Differential relations plus a diagram of the unit vectors for spherical coord<strong>in</strong>ates.


C.3. FRENET-SERRET COORDINATES 513<br />

The distance and volume elements, the cartesian coord<strong>in</strong>ate components of the spherical unit basis<br />

vectors, and the unit vector time derivatives are shown <strong>in</strong> the table given <strong>in</strong> figure 4. The time dependence<br />

of the unit vectors is used to derive the acceleration. As for the case of cyl<strong>in</strong>drical coord<strong>in</strong>ates, the ˆr ˆθ and<br />

ˆφ components of the acceleration <strong>in</strong>volve coupl<strong>in</strong>g of the coord<strong>in</strong>ates and their time derivatives.<br />

It is important to note that the angular unit vectors ˆθ and ˆφ are taken to be tangential to the circles of<br />

rotation. However, for discussion of angular velocity of angular momentum it is more convenient to use the<br />

axes of rotation def<strong>in</strong>ed by ˆr × ˆθ and ˆr × ˆφ for specify<strong>in</strong>g the vector properties which is perpendicular to<br />

the unit vectors ˆθ and ˆφ. Be careful not to confuse the unit vectors ˆθ and ˆφ with those used for the angular<br />

velocities ˙ and ˙.<br />

C.3 Frenet-Serret coord<strong>in</strong>ates<br />

The cartesian, polar, cyl<strong>in</strong>drical, or spherical curvil<strong>in</strong>ear coord<strong>in</strong>ate systems, all are orthogonal coord<strong>in</strong>ate<br />

systems that are fixed <strong>in</strong> space. There are situations where it is more convenient to use the Frenet-Serret<br />

coord<strong>in</strong>ates which comprise an orthogonal coord<strong>in</strong>ate system that is fixed to the particle that is mov<strong>in</strong>g<br />

along a cont<strong>in</strong>uous, differentiable, trajectory <strong>in</strong> three-dimensional Euclidean space. Let () represent a<br />

monotonically <strong>in</strong>creas<strong>in</strong>g arc-length along the trajectory of the particle motion as a function of time . The<br />

Frenet-Serret coord<strong>in</strong>ates, shown <strong>in</strong> figure 5 are the three <strong>in</strong>stantaneous orthogonal unit vectors ˆt ˆn and<br />

ˆb where the tangent unit vector ˆt is the <strong>in</strong>stantaneous tangent to the curve, the normal unit vector ˆn is <strong>in</strong><br />

the plane of curvature of the trajectory po<strong>in</strong>t<strong>in</strong>g towards the center of the <strong>in</strong>stantaneous radius of curvature<br />

and is perpendicular to the tangent unit vector ˆt while the b<strong>in</strong>ormal unit vector is ˆb =ˆt × ˆn which is the<br />

perpendicular to the plane of curvature and is mutually perpendicular to the other two Frenet-Serrat unit<br />

vectors. The Frenet-Serret unit vectors are def<strong>in</strong>ed by the relations<br />

ˆt<br />

<br />

ˆb<br />

<br />

ˆn<br />

<br />

= ˆn (C.16)<br />

= − ˆn (C.17)<br />

= −ˆt+ ˆb (C.18)<br />

The curvature = 1 <br />

where is the radius of curvature and is the torsion that can be either positive<br />

or negative. For <strong>in</strong>creas<strong>in</strong>g anon-zerocurvature implies that the triad of unit vectors rotate <strong>in</strong> a<br />

right-handed sense about ˆb. If the torsion is positive (negative) the triad of unit vectors rotates <strong>in</strong> right<br />

(left) handed sense about ˆt.<br />

Distance element<br />

Unit vectors<br />

Time derivatives<br />

of unit vectors<br />

<br />

<br />

⎛<br />

⎝<br />

ˆt<br />

ˆn<br />

ˆb<br />

¯<br />

s() =ˆt<br />

¯ r()<br />

<br />

ˆt() = v()<br />

|()|<br />

ˆn() = ˆt<br />

|ˆt|<br />

ˆb()=ˆt × ˆn<br />

¯ = ˆt()<br />

⎞ ⎛<br />

⎠ = || ⎝ 0 0 ⎞ ⎛<br />

− 0 ⎠ ⎝<br />

0 − 0<br />

ˆt<br />

ˆn<br />

ˆb<br />

⎞<br />

⎠<br />

b^<br />

n^<br />

^t<br />

Velocity<br />

Acceleration<br />

v() = r()<br />

<br />

a() = <br />

ˆt+ 2ˆn<br />

Table 5. The differential relations plus a diagram of the correspond<strong>in</strong>g unit vectors for the Frenet-Serret<br />

coord<strong>in</strong>ate system.


514 APPENDIX C. ORTHOGONAL COORDINATE SYSTEMS<br />

The above equations also can be rewritten <strong>in</strong> the form us<strong>in</strong>g a new unit rotation vector ω where<br />

ω=ˆt+ˆb<br />

(C.19)<br />

Then equations 16 − 18 are transformed to<br />

ˆt<br />

<br />

= ω × ˆt (C.20)<br />

ˆn<br />

<br />

= ω × ˆn (C.21)<br />

ˆb<br />

<br />

= ω × ˆb (C.22)<br />

In general the Frenet-Serret unit vectors are time dependent. If the curvature =0then the curve is a<br />

straight l<strong>in</strong>e and ˆn and ˆb are not well def<strong>in</strong>ed. If the torsion is zero then the trajectory lies <strong>in</strong> a plane. Note<br />

that a helix has constant curvature and constant torsion.<br />

Therateofchangeofageneralvectorfield E along the trajectory can be written as<br />

µ<br />

E<br />

= <br />

ˆt + <br />

<br />

ˆn+ <br />

ˆb + ω × E (C.23)<br />

The Frenet-Serret coord<strong>in</strong>ates are used <strong>in</strong> the life sciences to describe the motion of a mov<strong>in</strong>g organism<br />

<strong>in</strong> a viscous medium. The Frenet-Serret coord<strong>in</strong>ates also have applications to General Relativity.<br />

Problems<br />

1. The goal of this problem is to help you understand the orig<strong>in</strong> of the equations that relate two different coord<strong>in</strong>ate<br />

systems. Refer to diagrams for cyl<strong>in</strong>drical and spherical coord<strong>in</strong>ates as your teach<strong>in</strong>g assistant expla<strong>in</strong>s how to<br />

arrive at expressions for 1 2 and 3 <strong>in</strong> terms of and and how to derive expressions for the velocity and<br />

acceleration vectors <strong>in</strong> cyl<strong>in</strong>drical coord<strong>in</strong>ates. Now try to relate spherical and rectangular coord<strong>in</strong>ate systems.<br />

Your group should derive expressions relat<strong>in</strong>g the coord<strong>in</strong>ates of the two systems, expressions relat<strong>in</strong>g the unit<br />

vectors and their time derivatives of the two systems, and f<strong>in</strong>ally, expressions for the velocity and acceleration<br />

<strong>in</strong> spherical coord<strong>in</strong>ates.


Appendix D<br />

Coord<strong>in</strong>ate transformations<br />

Coord<strong>in</strong>ate systems can be translated, or rotated with respect to each other as well as be<strong>in</strong>g subject to spatial<br />

<strong>in</strong>version or time reversal. Scalars, vectors, and tensors are def<strong>in</strong>ed by their transformation properties under<br />

rotation, spatial <strong>in</strong>version and time reversal, and thus such transformations play a pivotal role <strong>in</strong> physics.<br />

D.1 Translational transformations<br />

Translational transformations are <strong>in</strong>volved frequently for transform<strong>in</strong>g between the center of mass and laboratory<br />

frames for reaction k<strong>in</strong>ematics as well as when perform<strong>in</strong>g vector addition of central forces for the<br />

cases where the centers are displaced. Both the classical Galilean transformation or the relativistic Lorentz<br />

transformation are handled the same way. Consider two parallel orthonormal coord<strong>in</strong>ate frames where the<br />

orig<strong>in</strong> of 0 ( 0 0 0 ) is displaced by a time dependent vector a() from the orig<strong>in</strong> of frame ( ). Then<br />

the Galilean transformation for a vector r <strong>in</strong> frame to r 0 <strong>in</strong> frame 0 is given by<br />

r ( 0 0 0 )=r ( )+a()<br />

(D.1)<br />

The velocities for a mov<strong>in</strong>g frame are given by the vector difference of the velocity <strong>in</strong> a stationary frame,<br />

and the velocity of the orig<strong>in</strong> of the mov<strong>in</strong>g frame. L<strong>in</strong>ear accelerations can be handled similarly.<br />

D.2 Rotational transformations<br />

D.2.1<br />

Rotation matrix<br />

Rotational transformations of the coord<strong>in</strong>ate system are used extensively <strong>in</strong> physics. The transformation<br />

properties of fields under rotation def<strong>in</strong>e the scalar and vector properties of fields, as well as rotational<br />

symmetry and conservation of angular momentum.<br />

Rotation of the coord<strong>in</strong>ate frame does not change the value of any scalar observable such as mass,<br />

temperature etc. That is, transformation of a scalar quantity is <strong>in</strong>variant under coord<strong>in</strong>ate rotation from<br />

→ 0 0 0 .<br />

( 0 0 0 )=()<br />

(D.2)<br />

By contrast, the components of a vector along the coord<strong>in</strong>ate axes change under rotation of the coord<strong>in</strong>ate<br />

axes. This difference <strong>in</strong> transformation properties under rotation between a scalar and a vector is important<br />

and def<strong>in</strong>es both scalars and a vectors.<br />

Matrix mechanics, described <strong>in</strong> appendix , provides the most convenient way to handle coord<strong>in</strong>ate<br />

rotations. The transformation matrix, between coord<strong>in</strong>ate systems hav<strong>in</strong>g differ<strong>in</strong>g orientations is called the<br />

rotation matrix. This transforms the components of any vector with respect to one coord<strong>in</strong>ate frame to<br />

the components with respect to a second coord<strong>in</strong>ate frame rotated with respect to the first frame.<br />

Assume a po<strong>in</strong>t has coord<strong>in</strong>ates ( 1 2 3 ) with respect to a certa<strong>in</strong> coord<strong>in</strong>ate system. Consider<br />

rotation to another coord<strong>in</strong>ate frame for which the po<strong>in</strong>t has coord<strong>in</strong>ates ( 0 1 0 2 0 3) and assume that the<br />

515


516 APPENDIX D. COORDINATE TRANSFORMATIONS<br />

orig<strong>in</strong>s of both frames co<strong>in</strong>cide. Rotation of a frame does not change the vector, only the vector components<br />

of the unit basis states. Therefore<br />

x = ê 0 1 0 1 + ê 0 2 0 2 + ê 0 3 0 3 = ê 1 1 + ê 2 2 + ê 3 3<br />

(D.3)<br />

Note that if one designates that the unit vectors for the unprimed coord<strong>in</strong>ate frame are (ê 1 ê 2 ê 3 ) and for<br />

the primed coord<strong>in</strong>ate frame (ê 0 1 ê 0 2 ê 0 3) then tak<strong>in</strong>g the scalar product of equation 3 sequentially with<br />

each of the unit base vectors (ê 0 1 ê 0 2 ê 0 3) leads to the follow<strong>in</strong>g three relations<br />

0 1 = (ê 0 1·ê 1 ) 1 +(ê 0 1·ê 2 ) 2 +(ê 0 1·ê 3 ) 3 (D.4)<br />

0 2 = (ê 0 2·ê 1 ) 1 +(ê 0 2·ê 2 ) 2 +(ê 0 2·ê 3 ) 3<br />

0 3 = (ê 0 3·ê 1 ) 1 +(ê 0 3·ê 2 ) 2 +(ê 0 3·ê 3 ) 3<br />

Note that the (ê 0 ·ê ) are the direction cos<strong>in</strong>es as def<strong>in</strong>ed by the scalar product of two unit vectors for axes<br />

, that is, they are the cos<strong>in</strong>e of the angle between the two unit vectors.<br />

Equation 4 can be written <strong>in</strong> matrix form as<br />

x 0 = λ · x<br />

where the “·” meansthe<strong>in</strong>ner matrix product of the rotation matrix λ and the vector x where<br />

⎛ ⎞<br />

⎛<br />

x 0 ≡ ⎝ 0 1<br />

0 ⎠<br />

2<br />

x ≡ ⎝ ⎞<br />

⎛<br />

⎞<br />

1<br />

2<br />

⎠ λ ≡ ⎝ ê0 1·ê 1 ê 0 1·ê 2 ê 0 1·ê 3<br />

ê 0 2·ê 1 ê 0 2·ê 2 ê 0 2·ê 3<br />

⎠<br />

0 3<br />

3<br />

ê 0 3·ê 1 ê 0 3·ê 2 ê 0 3·ê 3<br />

(D.5)<br />

(D.6)<br />

The <strong>in</strong>verse procedure is obta<strong>in</strong>ed by multiply<strong>in</strong>g equation 3 successively by one of the unit basis<br />

vectors (ê 1 ê 2 ê 3 ) lead<strong>in</strong>g to three equations<br />

Equation 7 can be written <strong>in</strong> matrix form as<br />

1 = (ê 1·ê 0 1) 0 1 +(ê 1·ê 0 2) 0 2 +(ê 1·ê 0 3) 0 3 (D.7)<br />

2 = (ê 2·ê 0 1) 0 1 +(ê 2·ê 0 2) 0 2 +(ê 2·ê 0 3) 0 3<br />

3 = (ê 3·ê 0 1) 0 1 +(ê 3·ê 0 2) 0 2 +(ê 3·ê 0 3) 0 3<br />

x = λ ·x 0<br />

(D.8)<br />

where λ is the transpose of λ.<br />

Note that substitut<strong>in</strong>g equation 5 <strong>in</strong>to equation 8 gives<br />

³ ´<br />

x = λ · (λ · x) = λ ·λ ·x (D.9)<br />

Thus ³ ´<br />

λ ·λ = I<br />

where I is the identity matrix. This implies that the rotation matrix λ is orthogonal with λ = λ −1 .<br />

It is convenient to rename the elements of the rotation matrix to be<br />

≡ (ê 0 ·ê )<br />

(D.10)<br />

so that the rotation matrix is written more compactly as<br />

⎛<br />

λ ≡ ⎝ ⎞<br />

11 12 13<br />

21 22 23<br />

⎠<br />

31 32 33<br />

and equation 4 becomes<br />

0 1 = 11 1 + 12 2 + 13 3 (D.11)<br />

0 2 = 21 1 + 22 2 + 23 3<br />

0 3 = 31 1 + 32 2 + 33 3


D.2. ROTATIONAL TRANSFORMATIONS 517<br />

Consider an arbitrary rotation through an angle . Equations (10) and (11) can be used to relate<br />

six of the n<strong>in</strong>e quantities <strong>in</strong> the rotation matrix, so only three of the quantities are <strong>in</strong>dependent. That<br />

is, because of equation (11) we have three equations which ensure that the transformation is unitary.<br />

2 1 + 2 2 + 2 3 =1<br />

(D.12)<br />

Also requir<strong>in</strong>g that the axes be orthogonal gives three equations<br />

X<br />

=0 6= (D.13)<br />

These six relations can be expressed as<br />

<br />

X<br />

= <br />

<br />

(D.14)<br />

The fact that the rotation matrix should have three <strong>in</strong>dependent quantities is due to the fact that all rotations<br />

can be expressed <strong>in</strong> terms of rotations about three orthogonal axes.<br />

D.1 Example: Rotation matrix:<br />

Consider a po<strong>in</strong>t ( 1 2 3 )= (3 4 5) <strong>in</strong> the unprimed coord<strong>in</strong>ate system. Consider the same po<strong>in</strong>t<br />

( 0 1 0 2 0 3) <strong>in</strong> the primed coord<strong>in</strong>ate system which has been rotated by an angle 60 ◦ about the 1 axis as<br />

shown. The direction cos<strong>in</strong>es 0 =cos( 0 ) can be determ<strong>in</strong>ed from the figure to be the follow<strong>in</strong>g<br />

0 0 0 =cos( 0 )<br />

1 1 0 1<br />

1 2 90 0<br />

1 3 90 0<br />

2 1 90 0<br />

2 2 60 0500<br />

2 3 90 − 60 0866<br />

3 1 90 0<br />

3 2 90 + 60 −0866<br />

3 3 60 0500<br />

Thus the rotation matrix is<br />

⎛<br />

= ⎝ 1 0 0 ⎞<br />

0 0500 0866 ⎠<br />

0 −0866 0500<br />

The transform po<strong>in</strong>t P 0 (x 0 1 x 0 2 x 0 3) therefore is given by<br />

⎛ ⎞ ⎛<br />

⎝ 0 1<br />

0 ⎠<br />

2 = ⎝ 1 0 0 ⎞ ⎛<br />

0 0500 0866 ⎠ · ⎝ 3 4<br />

0 3 0 −0866 0500 5<br />

⎞ ⎛<br />

⎠ = ⎝<br />

3<br />

6330<br />

−0964<br />

Note that the radial coord<strong>in</strong>ate r = r 0 =√ 50. That is, the rotational transformation is unitary and thus<br />

the magnitude of the vector is unchanged.<br />

⎞<br />


518 APPENDIX D. COORDINATE TRANSFORMATIONS<br />

D.2 Example: Proof that a rotation matrix is orthogonal<br />

Consider the rotation matrix<br />

The product<br />

⎛<br />

λ ·λ = 1 ⎝ 4 1 8<br />

7 4 −4<br />

81<br />

−4 8 1<br />

which implies that is orthogonal.<br />

D.2.2<br />

F<strong>in</strong>ite rotations<br />

⎛<br />

λ = 1 ⎝ 4 7 −4 ⎞<br />

1 4 8 ⎠<br />

9<br />

8 −4 1<br />

⎞ ⎛<br />

⎠ ·<br />

⎝ 4 7 −4<br />

1 4 8<br />

8 −4 1<br />

⎞ ⎛<br />

⎠ = 1 ⎝ 81 0 0<br />

0 81 0<br />

81<br />

0 0 81<br />

⎞<br />

⎠ =1<br />

Consider two f<strong>in</strong>ite 90 rotations and<br />

illustrated <strong>in</strong> figure 1 The rotation<br />

is 90 around the 3 axis <strong>in</strong> a<br />

right-handed direction as shown. In such<br />

A B<br />

a rotation the axes transform to 0 1 = 2 ,<br />

0 2 = − 1 , 0 3 = 3 and the rotation matrix<br />

= 90° = 90°<br />

3 2<br />

is ⎛<br />

x 3<br />

λ = ⎝ 0 1 0 ⎞<br />

−1 0 0 ⎠ (D.15)<br />

0 0 1<br />

The second rotation λ is a right-handed<br />

rotation about the 0 1 axis which formerly<br />

was the 2 axis. Then ” 1 = 0 2, ” 2 = − 0 1,<br />

x 1<br />

x 2<br />

B<br />

A<br />

” 3 = 0 3 and the rotation matrix is<br />

⎛<br />

= 90° = 90°<br />

2 3<br />

λ = ⎝ 1 0 0 ⎞<br />

0 0 1 ⎠ (D.16)<br />

0 −1 0<br />

Consider the product of these two f<strong>in</strong>ite rotations<br />

Figure D.1: Order of two f<strong>in</strong>ite rotations for a parallelepiped.<br />

which corresponds to a s<strong>in</strong>gle rota-<br />

tion matrix λ <br />

λ = λ λ (D.17)<br />

That is:<br />

⎛<br />

λ = ⎝ 1 0 0 ⎞ ⎛<br />

0 0 1 ⎠ ⎝ 0 1 0 ⎞ ⎛<br />

−1 0 0 ⎠ = ⎝ 0 1 0 ⎞<br />

0 0 1 ⎠<br />

0 −1 0 0 0 1 1 0 0<br />

Now consider that the order of these two rotations is reversed.<br />

λ = λ λ <br />

(D.18)<br />

(D.19)<br />

That is:<br />

⎛<br />

λ = ⎝ 0 1 0 ⎞ ⎛<br />

−1 0 0 ⎠ ⎝ 1 0 0 ⎞ ⎛<br />

0 0 1 ⎠ = ⎝ 0 0 1 ⎞<br />

−1 0 0 ⎠ 6= λ <br />

(D.20)<br />

0 0 1 0 −1 0 0 −1 0<br />

An entirely different orientation results as illustrated <strong>in</strong> figure 1.<br />

This behavior of f<strong>in</strong>ite rotations is a consequence of the fact that f<strong>in</strong>ite rotations do not commute, that<br />

is, revers<strong>in</strong>g the order does not give the same answer. Thus, if we associate the vectors A and B with<br />

these rotations, then it implies that the vector product AB 6= BA. Thatis,forf<strong>in</strong>ite rotation matrices, the<br />

product does not behave like for true vectors s<strong>in</strong>ce they do not commute.


D.2. ROTATIONAL TRANSFORMATIONS 519<br />

D.2.3<br />

Inf<strong>in</strong>itessimal rotations<br />

Inf<strong>in</strong>itessimal rotations do not suffer from the noncommutation defect<br />

of f<strong>in</strong>ite rotations. If the position vector of a po<strong>in</strong>t changes from r to<br />

r + r then the geometrical situation is represented correctly by<br />

r = θ × r<br />

(D.21)<br />

where θ is a quantity whose magnitude is equal to the <strong>in</strong>f<strong>in</strong>itessimal<br />

rotation angle and which has a direction along the <strong>in</strong>stantaneous axis<br />

of rotation as illustrated <strong>in</strong> figure 2.<br />

The <strong>in</strong>f<strong>in</strong>itessimal angle θ is a vector which is shown by prov<strong>in</strong>g<br />

that two <strong>in</strong>f<strong>in</strong>itessimal rotations θ 1 and θ 2 commute. The change<br />

<strong>in</strong> position vectors of the po<strong>in</strong>t are<br />

r 1 = θ 1 × r<br />

and<br />

r 2 = θ 2 × (r + r 1 )<br />

Thus the f<strong>in</strong>al position vector for θ 1 followed by θ 2 is<br />

r + r 1 + r 2 = r + θ 1 × r + θ 2 × (r + r 1 )<br />

(D.22)<br />

(D.23)<br />

(D.24)<br />

Figure D.2: Inf<strong>in</strong>itessimal rotation<br />

Assum<strong>in</strong>g that the second-order <strong>in</strong>f<strong>in</strong>itessimals can be ignored gives<br />

r + r 1 + r 2 = r + θ 1 × r + θ 2 × r<br />

(D.25)<br />

Consider now the <strong>in</strong>verse order of rotations.<br />

r + r 2 + r 1 = r + θ 2 × r + θ 1 × (r + r 2 )<br />

(D.26)<br />

Aga<strong>in</strong>, neglect<strong>in</strong>g the second-order <strong>in</strong>f<strong>in</strong>itessimals gives<br />

r + r 2 + r 1 = r + θ 2 × r + θ 1 × r<br />

(D.27)<br />

Note that the products of these two <strong>in</strong>f<strong>in</strong>itessimal rotations, 25 and 27 are identical. That is, assum<strong>in</strong>g<br />

that second-order <strong>in</strong>f<strong>in</strong>itessimals can be neglected, then the <strong>in</strong>f<strong>in</strong>itessimal rotations commute, and thus θ 1<br />

and θ 2 are correctly represented by vectors.<br />

The fact that θ is a vector allows angular velocity to be represented by a vector. That is, angular<br />

velocity is the ratio of an <strong>in</strong>f<strong>in</strong>itessimal rotation to an <strong>in</strong>f<strong>in</strong>itessimal time.<br />

ω = θ<br />

(D.28)<br />

<br />

Note that this implies that the velocity of the po<strong>in</strong>t can be expressed as<br />

v = r<br />

= θ<br />

× r = ω × r<br />

(D.29)<br />

D.2.4<br />

Proper and improper rotations<br />

The requirement that the coord<strong>in</strong>ate axes be orthogonal, and that the transformation be unitary, leads to<br />

the relation between the components of the rotation matrix.<br />

X<br />

= <br />

(D.30)<br />

<br />

It was shown <strong>in</strong> equation 12 that, for such an orthogonal matrix, the <strong>in</strong>verse matrix −1 equals the<br />

transposed matrix <br />

λ −1 = λ


520 APPENDIX D. COORDINATE TRANSFORMATIONS<br />

Insert<strong>in</strong>g the orthogonality relation for the rotation matrix leads to the fact that the square of the determ<strong>in</strong>ant<br />

of the rotation matrix equals one,<br />

|| 2 =1<br />

(D.31)<br />

that is<br />

|| = ±1<br />

(D.32)<br />

A proper rotation is the rotation of a normal vector and has<br />

|| =+1<br />

(D.33)<br />

An improper rotation corresponds to<br />

|| = −1<br />

(D.34)<br />

An improper rotation implies a rotation plus a spatial reflection which cannot be achieved by any comb<strong>in</strong>ation<br />

of only rotations.<br />

Consider the cross product of two vectors c = a × b It can be shown that the cross product behaves<br />

under rotation as:<br />

0 = || X <br />

(D.35)<br />

<br />

For all proper rotations the determ<strong>in</strong>ant of =+1and thus the cross product also acts like a proper vector<br />

under rotation. This is not true for improper rotations where || = −1<br />

D.3 Spatial <strong>in</strong>version transformation<br />

Spatial <strong>in</strong>version, that is, mirror reflection, corresponds to reflection of all coord<strong>in</strong>ate vectors, b i = − b i b j = −<br />

b j and k b = − k b Such a transformation corresponds to the transformation matrix<br />

⎛<br />

⎞ ⎛ ⎞<br />

λ =<br />

⎝ −1 0 0<br />

0 −1 0<br />

0 0 −1<br />

Thus || = −1 that is, it corresponds to an improper<br />

rotation. A spatial <strong>in</strong>version for two vectors A() and<br />

B() correspond to<br />

⎠ = −<br />

⎝ 1 0 0<br />

0 1 0<br />

0 0 1<br />

⎠<br />

x 3<br />

x 2<br />

x 3<br />

x 1 ‘<br />

(D.36)<br />

A() = −A(−) (D.37)<br />

B() = −B(−)<br />

That is, normal polar vectors change sign under spatial<br />

reflection. However, the cross product C = A × B<br />

does not change sign under spatial <strong>in</strong>version s<strong>in</strong>ce the<br />

product of the two m<strong>in</strong>us signs is positive. That is,<br />

x 1<br />

x 2<br />

‘<br />

‘<br />

C() =+C(−)<br />

(D.38)<br />

Figure D.3: Inversion of an object corresponds to<br />

Thus the cross product behaves differently from a polar reflection about the orig<strong>in</strong> of all axes.<br />

vector. This improper behavior is characteristic of an<br />

axial vector, which also is called a pseudovector.<br />

Examples of pseudovectors are angular momentum, sp<strong>in</strong>, magnetic field etc. These pseudovectors are<br />

def<strong>in</strong>ed us<strong>in</strong>g the right-hand rule and thus have handedness. For a right-handed system<br />

C = A × B<br />

(D.39)<br />

Chang<strong>in</strong>g to a left-handed system leads to<br />

C = B × A = −A × B<br />

(D.40)


D.4. TIME REVERSAL TRANSFORMATION 521<br />

That is, handedness corresponds to a def<strong>in</strong>ite order<strong>in</strong>g of the cross product. Proper orthogonal transformations<br />

are said to preserve chirality (Greek for handedness) of a coord<strong>in</strong>ate system.<br />

An example of the use of the right-handed system is the usual def<strong>in</strong>ition of cartesian unit vectors,<br />

b i × b j = b k<br />

(D.41)<br />

An obvious question to be asked, is the handedness of a coord<strong>in</strong>ate system merely a mathematical curiosity<br />

or does it have some deep underly<strong>in</strong>g significance? Consider the Lorentz force<br />

F = (E + v × B)<br />

(D.42)<br />

S<strong>in</strong>ce force and velocity are proper vectors then the magnetic B field must be a pseudo vector. Note that<br />

calculation of the B field occurs only <strong>in</strong> cross products such as,<br />

∇ × B = j<br />

(D.43)<br />

where the current density j is a proper vector. Another example is the Biot-Savart Law which expresses B<br />

as<br />

B = l × r<br />

4 2<br />

(D.44)<br />

Thus even though B is a pseudo vector, the force F rema<strong>in</strong>s a proper vector. Thus if a left-handed coord<strong>in</strong>ate<br />

def<strong>in</strong>ition of B = r×l<br />

4 <br />

is used <strong>in</strong> 44, andF = (E + B 2<br />

×v) <strong>in</strong> 42 then the same f<strong>in</strong>al physical<br />

result would be obta<strong>in</strong>ed.<br />

It was long thought that the laws of physics were symmetric with respect to spatial <strong>in</strong>version ( i.e. mirror<br />

reflection), mean<strong>in</strong>g that the choice between a left-handed and right-handed representations (chirality) was<br />

arbitrary. This is true for gravitational, electromagnetic and the strong force, and is called the conservation<br />

of parity. The fourth fundamental force <strong>in</strong> nature, the weak force, violates parity and favours handedness.<br />

It turns out that right-handed ord<strong>in</strong>ary matter is symmetrical with left-handed antimatter.<br />

In addition to the two flavours of vectors, one has scalars and pseudoscalars def<strong>in</strong>ed by:<br />

An example of a pseudoscalar is the scalar product A · (B × C)<br />

() = + (−) (D.45)<br />

() = − (−) (D.46)<br />

D.4 Time reversal transformation<br />

The basic laws of classical mechanics are <strong>in</strong>variant to the sense of the direction of time. Under time reversal<br />

the vector r is unchanged while both momentum p and time change sign under time reversal, thus the time<br />

derivative F = p<br />

p<br />

<br />

is <strong>in</strong>variant to time reversal; that is, the force is unchanged and Newton’s Laws F =<br />

<br />

are <strong>in</strong>variant under time reversal. S<strong>in</strong>ce the force can be expressed as the gradient of a scalar potential for<br />

a conservative field, then the potential also rema<strong>in</strong>s unchanged. That is<br />

p<br />

= −∇() =F<br />

(D.47)<br />

It is necessary to <strong>in</strong>troduce tensor algebra, given <strong>in</strong> appendix , prior to discussion of the transformation<br />

properties of observables which is the topic of appendix 5.<br />

Problems<br />

1. Suppose the 2 -axis of a rectangular coord<strong>in</strong>ate system is rotated by 30 ◦ away from the 3 -axis around the<br />

1 -axis.<br />

(a) F<strong>in</strong>d the correspond<strong>in</strong>g transformation matrix. Try to do this by draw<strong>in</strong>g a diagram <strong>in</strong>stead of go<strong>in</strong>g to<br />

the book or the notes for a formula.


522 APPENDIX D. COORDINATE TRANSFORMATIONS<br />

(b) Is this an orthogonal matrix? If so, show that it satisfies the ma<strong>in</strong> properties of an orthogonal matrix. If<br />

not, expla<strong>in</strong> why it fails to be orthogonal.<br />

(c) Does this matrix represent a proper or an improper rotation? How do you know?<br />

2. When you were first <strong>in</strong>troduced to vectors, you most likely were told that a scalar is a quantity that is def<strong>in</strong>ed<br />

by a magnitude, while a vector has both a magnitude and a direction. While this is certa<strong>in</strong>ly true, there is<br />

another, more sophisticated way to def<strong>in</strong>e a scalar quantity and a vector quantity: through their transformation<br />

properties. A scalar quantity transforms as 0 = while a vector quantity transforms as 0 = P To<br />

show that the scalar product does <strong>in</strong>deed transform as a scalar, note that:<br />

⎛ ⎞<br />

A 0·B 0 = X 0 0 = X ⎝ X Ã ! X<br />

<br />

⎠ = X Ã ! X<br />

<br />

<br />

<br />

<br />

<br />

= X <br />

à X<br />

<br />

!<br />

= X <br />

= A · B<br />

Now you will show that the vector product transforms as a vector. Beg<strong>in</strong> by writ<strong>in</strong>g out what you are try<strong>in</strong>g<br />

to show explicitly and show it to the teach<strong>in</strong>g assistant. Once the teach<strong>in</strong>g assistant has confirmed that you<br />

have the correct expression, try to prove it. The vector product is a bit more difficult to work with than the<br />

scalar product, so your teach<strong>in</strong>g assistant is prepared to give you a h<strong>in</strong>t if you get stuck.<br />

3. Suppose you have two rectangular coord<strong>in</strong>ate systems that share a common orig<strong>in</strong>, but one system is rotated<br />

by an angle with respect to the other. To describe this rotation, you have made use of the rotation matrix<br />

(). (I’m chang<strong>in</strong>g the notation slightly to put the emphasis on the angle of rotation.)<br />

(a) Verify that the product of two rotation matrices ( 1 )( 2 ) is <strong>in</strong> itself a rotation matrix.<br />

(b) In abstract algebra, a group is def<strong>in</strong>ed as a set of elements together with a b<strong>in</strong>ary operation ∗ act<strong>in</strong>g<br />

on that set such that four properties are satisfied:<br />

i. (Closure) For any two elements and <strong>in</strong> the group , the product of the elements, ∗ is also<br />

<strong>in</strong> the group .<br />

ii. (Associativity) For any three elements of the group , ( ∗ ) ∗ = ∗ ( ∗ ).<br />

iii. (Existence of Identity) The group conta<strong>in</strong>s an identity element such that ∗ = ∗ = for<br />

all ∈ .<br />

iv. (Existence of Inverses) For each element ∈ , there exists an <strong>in</strong>verse element −1 ∈ such that<br />

∗ −1 = −1 ∗ = .<br />

Show that if the product ∗ denotes the product of two matrices, then the set of rotation matrices together<br />

with ∗ forms a group. This group is known as the special orthogonal group <strong>in</strong> two dimensions, also known<br />

as (2).<br />

(c) Is this group commutative? In abstract algebra, a commutative group is called an abelian group.<br />

4. When you look <strong>in</strong> a mirror the image of you appears left-to-right reversed, that is, the image of your left ear<br />

appears to be the right ear of the image and vise versa. Expla<strong>in</strong> why the image is left-right reversed rather<br />

than up-down reversed or reversed about some other axis; i.e. expla<strong>in</strong> what breaks the symmetry that leads to<br />

these properties of the mirror image.<br />

5. F<strong>in</strong>d the transformation matrix that rotates the axis 3 of a rectangular coord<strong>in</strong>ate system 45 toward 1<br />

around the 2 axis.<br />

6. For simplicity, take to be a two-dimensional transformation matrix. Show by direct expansion that |λ| 2 =1.


Appendix E<br />

Tensor algebra<br />

E.1 Tensors<br />

Mathematically scalars and vectors are the first two members of a hierarchy of entities, called tensors,<br />

that behave under coord<strong>in</strong>ate transformations as described <strong>in</strong> appendix . The use of the tensor notation<br />

provides a compact and elegant way to handle transformations <strong>in</strong> physics.<br />

Ascalarisarank0 tensor with one component, that is <strong>in</strong>variant under change of the coord<strong>in</strong>ate system.<br />

( 0 0 0 )=()<br />

(E.1)<br />

A vector is a rank 1 tensor which has three components, that transform under rotation accord<strong>in</strong>g to<br />

matrix relation<br />

x 0 = λ · x<br />

(E.2)<br />

where λ is the rotation matrix. Equation 2 can be written <strong>in</strong> the suffix formas<br />

3X<br />

0 = <br />

(E.3)<br />

=1<br />

The above def<strong>in</strong>itions of scalars and vectors can be subsumed <strong>in</strong>to a class of entities called tensors of rank <br />

that have 3 components. A scalar is a tensor of rank =0,withonly3 0 =1component, whereas a vector<br />

has rank =1 that is, the vector x has one suffix and 3 1 =3components.<br />

A second-order tensor has rank =2with two suffixes, that is, it has 3 2 =9components that<br />

transform under rotation as<br />

3X 3X<br />

0 = <br />

(E.4)<br />

=1 =1<br />

For second-order tensors, the transformation formula given by equation 4 can be written more compactly<br />

us<strong>in</strong>g matrices. Thus the second-order tensor can be written as a 3 × 3 matrix<br />

⎛<br />

T ≡ ⎝ ⎞<br />

11 12 13<br />

21 22 23<br />

⎠<br />

(E.5)<br />

31 32 33<br />

The rotational transformation given <strong>in</strong> equation 4 can be written <strong>in</strong> the form<br />

Ã<br />

3X 3X<br />

! Ã<br />

3X 3X<br />

!<br />

0 = = (E.6)<br />

=1 =1<br />

=1 =1<br />

where are the matrix elements of the transposed matrix λ . The summations <strong>in</strong> 6 can be expressed<br />

<strong>in</strong> both the tensor and conventional matrix form as the matrix product<br />

T 0 = λ · T · λ <br />

Equation 7 def<strong>in</strong>es the rotational properties of a spherical tensor.<br />

523<br />

(E.7)


524 APPENDIX E. TENSOR ALGEBRA<br />

E.2 Tensor products<br />

E.2.1<br />

Tensor outer product<br />

Tensor products feature prom<strong>in</strong>ently when us<strong>in</strong>g tensors to represent transformations. A second-order tensor<br />

T can be formed by us<strong>in</strong>g the tensor product, also called outer product, of two vectors a and b which,<br />

written<strong>in</strong>suffix form,is<br />

⎛<br />

T ≡ a ⊗ b = ⎝ ⎞<br />

1 1 1 2 1 3<br />

2 1 2 2 2 3<br />

⎠<br />

(E.8)<br />

3 1 3 2 3 3<br />

In component form the matrix elements of this matrix are given by<br />

= <br />

(E.9)<br />

This second-order tensor product has a rank =2 that is, it equals the sum of the ranks of the two<br />

vectors. Equation 8 is called a dyad s<strong>in</strong>ce it was derived by tak<strong>in</strong>g the dyadic product of two vectors. In<br />

general, multiplication, or division, of two vectors leads to second-order tensors. Note that this second-order<br />

tensor product completes the triad of tensors possible tak<strong>in</strong>g the product of two vectors. That is, the scalar<br />

product a · b, hasrank =0, the vector product a × b, rank =1and the tensor product a ⊗ b has rank 1<br />

=2.<br />

Higher-order tensors can be created by tak<strong>in</strong>g more complicated tensor products. For example, a rank-3<br />

tensor can be created by tak<strong>in</strong>g the tensor outer product of the rank-2 tensor and a vector which, for<br />

a dyadic tensor, can be written as the tensor product of three vectors. That is,<br />

= = <br />

(E.10)<br />

In summary, the rank of the tensor product equals the sum of the ranks of the tensors <strong>in</strong>cluded <strong>in</strong> the tensor<br />

product.<br />

E.2.2<br />

Tensor <strong>in</strong>ner product<br />

The lowest rank tensor product, which is called the <strong>in</strong>ner product, is obta<strong>in</strong>ed by tak<strong>in</strong>g the tensor product<br />

of two tensors for the special case where one <strong>in</strong>dex is repeated, and tak<strong>in</strong>g the sum over this repeated <strong>in</strong>dex.<br />

Summ<strong>in</strong>g over this repeated <strong>in</strong>dex, which is called contraction, removes the two <strong>in</strong>dices for which the <strong>in</strong>dex<br />

is repeated, result<strong>in</strong>g <strong>in</strong> a tensor that has rank equal to the sum of the ranks m<strong>in</strong>us 2 for one contraction.<br />

That is, the product tensor has rank = 1 + 2 − 2.<br />

The simplest example is the <strong>in</strong>ner product of two vectors which has rank =1+1− 2=0,thatis,itis<br />

the scalar product that equals the trace of the <strong>in</strong>ner product matrix, and this <strong>in</strong>ner product is commutative.<br />

An especially important case is the <strong>in</strong>ner product of a rank-2 dyad a ⊗ b given by equation 8 with a<br />

vector c, that is, the <strong>in</strong>ner product T = a ⊗ b · c. Written <strong>in</strong> component form, the <strong>in</strong>ner product is<br />

Ã<br />

3X<br />

3X<br />

!<br />

= =(a · b) (E.11)<br />

<br />

<br />

The scalar product a · b is a scalar number, and thus the <strong>in</strong>ner-product tensor is the vector c renormalized<br />

by the magnitude of the scalar product a · b. That is, it has a rank =2+1−2 =1. Thus the <strong>in</strong>ner product<br />

of this rank-2 tensor with a vector gives a vector. The <strong>in</strong>ner product of a rank-2 tensor with a rank-1 tensor<br />

is used <strong>in</strong> this book for handl<strong>in</strong>g the rotation matrix, the <strong>in</strong>ertia tensor for rigid-body rotation, and for the<br />

stress and the stra<strong>in</strong> tensors used to describe elasticity <strong>in</strong> solids.<br />

E.1 Example: Displacement gradient tensor<br />

The displacement gradient tensor provides an example of the use of the matrix representation to manipulate<br />

tensors. Let φ( 1 2 3 ) be a vector field expressed <strong>in</strong> a cartesian basis. The def<strong>in</strong>ition of the gradient<br />

G = ∇φ gives that<br />

φ = G·x<br />

1 The common convention is to denote the scalar product as a · b the vector product as a × b, and tensor product as a ⊗ b.


E.3. TENSOR PROPERTIES 525<br />

Calculat<strong>in</strong>g the components of φ <strong>in</strong> terms of x gives<br />

Us<strong>in</strong>g <strong>in</strong>dex notation this can be written as<br />

1 = 1<br />

1<br />

1 + 1<br />

2<br />

2 + 1<br />

3<br />

3<br />

2 = 2<br />

1<br />

1 + 2<br />

2<br />

2 + 2<br />

3<br />

3<br />

3 = 3<br />

1<br />

1 + 3<br />

2<br />

2 + 3<br />

3<br />

3<br />

= <br />

<br />

<br />

The second-rank gradient tensor G can be represented <strong>in</strong> the matrix form as<br />

1 1 1<br />

1 2 3<br />

<br />

G =<br />

2 2 2<br />

1 2 3<br />

¯<br />

¯¯¯¯¯¯¯<br />

3 3 3<br />

1 2 3<br />

Then the vector φ can be expressed compactly as the <strong>in</strong>ner product of G and xthat is<br />

E.3 Tensor properties<br />

φ = G·x<br />

In pr<strong>in</strong>ciple one must dist<strong>in</strong>guish between a 3×3 square matrix, and the tensor component representations of<br />

a rank-2 tensor. However, as illustrated by the previous discussion, for orthogonal transformations, the tensor<br />

components of the second rank tensor transform identically with the matrix components. Thus functionally,<br />

the matrix formulation and tensor representations are identical. As a consequence, all the term<strong>in</strong>ology and<br />

operations used <strong>in</strong> matrix mechanics are equally applicable to the tensor representation.<br />

The tensor representation of the rotation matrix provides the simplest example of the equivalence of<br />

the matrix and tensor representations of transformations. Appendix 2 showed that the unitary rotation<br />

matrix λ act<strong>in</strong>g on a vector x transforms it to the vector x 0 that is rotated with respect to x. Thatis,the<br />

transformation is<br />

x 0 = λ · x<br />

(5)<br />

where<br />

⎛ ⎞<br />

⎛ ⎞<br />

⎛<br />

⎞<br />

x 0 ≡<br />

⎝ 0 1<br />

0 2<br />

0 3<br />

⎠<br />

x ≡<br />

⎝ 1<br />

2<br />

3<br />

⎠<br />

λ ≡<br />

⎝ ê0 1·ê 1<br />

ê 0 2·ê 1<br />

ê 0 1·ê 2<br />

ê 0 2·ê 2<br />

ê 0 1·ê 3<br />

ê 0 2·ê 3<br />

ê 0 3·ê 1 ê 0 3·ê 2 ê 0 3·ê 3<br />

Appendix 2 showed that the rotation matrix λ requires 9 components to fully specify the transformation<br />

from the <strong>in</strong>itial 3-component vector x to the rotated vector x 0 . The rotation tensor is a dyad as well as be<strong>in</strong>g<br />

unitary and dimensionless. Note that equation 5 is an example of the <strong>in</strong>ner product of a rank−2 rotation<br />

tensor act<strong>in</strong>g on a vector lead<strong>in</strong>g to a another vectorthatisrotatedwithrespecttothefirst vector.<br />

In general, rank-2 tensors have dimensions and are not unitary. For example, the angular velocity vector<br />

ω and the angular momentum vector L are related by the <strong>in</strong>ner product of the <strong>in</strong>ertia tensor {I} and ω.<br />

That is<br />

L ={I}·ω<br />

(116)<br />

The <strong>in</strong>ertia tensor has dimensions of × 2 and relates two very different vector observables. The<br />

stress tensor and the stra<strong>in</strong> tensor, discussed <strong>in</strong> chapter 15 provide another example of second-order tensors<br />

that are used to transform one vector observable to another vector observable analogous to the case of the<br />

rotation matrix or the <strong>in</strong>ertia tensor.<br />

Note that pseudo-tensors can be used to make a rotational transformation plus a change <strong>in</strong> the sign.<br />

That is, they lead to a parity <strong>in</strong>version.<br />

The tensor notation is used extensively <strong>in</strong> physics s<strong>in</strong>ce it provides a powerful, elegant, and compact<br />

representation for describ<strong>in</strong>g transformations.<br />

⎠<br />

(6)


526 APPENDIX E. TENSOR ALGEBRA<br />

E.4 Contravariant and covariant tensors<br />

In general the configuration space used to specify a dynamical system is not a Euclidean space <strong>in</strong> that<br />

there may not be a system of coord<strong>in</strong>ates for which the distance between any two neighbor<strong>in</strong>g po<strong>in</strong>ts can<br />

be represented by the sum of the squares of the coord<strong>in</strong>ate differentials. For example, a set of cartesian<br />

coord<strong>in</strong>ate does not exist for the two-dimension motion of a s<strong>in</strong>gle particle constra<strong>in</strong>ed to the curved surface<br />

of a fixed sphere. Such curved spaces need to be represented <strong>in</strong> terms of Riemannian geometry rather<br />

than Euclidean geometry. Curved configuration spaces occur <strong>in</strong> some branches of physics such as E<strong>in</strong>ste<strong>in</strong>’s<br />

General Theory of Relativity.<br />

Tensors have transformation properties that can be either contravariant or covariant. Consider a set of<br />

generalized coord<strong>in</strong>ates 0 that are a function of the coord<strong>in</strong>ates . Then<strong>in</strong>f<strong>in</strong>itessimal changes will lead<br />

to <strong>in</strong>f<strong>in</strong>itessimal changes 0 where<br />

0 = X 0<br />

<br />

(E.12)<br />

<br />

Contravariant components of a tensor transform accord<strong>in</strong>g to the relation<br />

0 = X <br />

0<br />

<br />

(E.13)<br />

Equation 13 relates the contravariant components <strong>in</strong> the unprimed and primed frames.<br />

Derivatives of a scalar function , suchas<br />

0 = <br />

= X <br />

<br />

= X <br />

<br />

<br />

(E.14)<br />

That is, covariant components of the tensor transform accord<strong>in</strong>g to the relation<br />

0 = X <br />

<br />

<br />

(E.15)<br />

It is important to differentiate between contravariant and covariant vectors. The superscript/subscript<br />

convention for dist<strong>in</strong>guish<strong>in</strong>g between these two flavours of tensors is given <strong>in</strong> table 1<br />

Table 1. E<strong>in</strong>ste<strong>in</strong> notation for tensors.<br />

<br />

<br />

denotes a contravariant vector<br />

denotes a covariant vector<br />

In l<strong>in</strong>ear algebra one can map from one coord<strong>in</strong>ate system to another as illustrated <strong>in</strong> appendix . That<br />

is, the tensor x can be expressed as components with respect to either the unprimed or primed coord<strong>in</strong>ate<br />

frames<br />

x = ê 0 1 0 1 + ê 0 2 0 2 + ê 0 3 0 3 = ê 1 1 + ê 2 2 + ê 3 3<br />

(E.16)<br />

For a −dimensional manifold the unit basis column vectors ê transform accord<strong>in</strong>g to the transformation<br />

matrix λ<br />

ê 0 = λ · ê<br />

(E.17)<br />

S<strong>in</strong>ce the tensor x is <strong>in</strong>dependent of the coord<strong>in</strong>ate basis, the components of x must have the opposite<br />

transform<br />

x 0 = ¡ λ −1¢ <br />

·x<br />

(E.18)<br />

This normal vector x is called a “contravariant vector” because it transforms contrary to the basis column<br />

vector transformation.<br />

The <strong>in</strong>verse of equation 18 gives that the column vector element<br />

= X <br />

λ 0 <br />

(E.19)


E.5. GENERALIZED INNER PRODUCT 527<br />

Consider the case of a gradient with respect to the coord<strong>in</strong>ate x <strong>in</strong> both the unprimed and primed bases.<br />

Us<strong>in</strong>g the cha<strong>in</strong> rule for the partial derivative then the component of the gradient <strong>in</strong> the primed frame can<br />

be expanded as<br />

(∇) 0 = <br />

0 = X<br />

<br />

That is, the gradient transforms as<br />

<br />

0 = X<br />

<br />

∇ 0 = λ · ∇<br />

<br />

<br />

λ = <br />

<br />

(E.20)<br />

(E.21)<br />

That is, a gradient transforms as a covariant vector, like the unit vectors, whereas a vector is contravariant<br />

under transformation.<br />

Normally the basis is orthonormal, ¡ λ −1¢ <br />

= λ and thus there is no difference between contravariant and<br />

covariant vectors. However, for curved coord<strong>in</strong>ate systems, such as non-Euclidean geometry <strong>in</strong> the General<br />

Theory of Relativity, the covariant and contravariant vectors behave differently.<br />

The E<strong>in</strong>ste<strong>in</strong> convention is extended to apply to matrices by writ<strong>in</strong>g the elements of the matrix A as<br />

while the elements of the transposed matrix A −1 are written as . The matrix product for A with a<br />

contravariant vector X is written as<br />

0 = X <br />

(E.22)<br />

<br />

where the summation over effectively cancels the identical superscript and subscript .<br />

Similarly a covariant vector, such as a gradient, is written as,<br />

¡<br />

∇ 0 ¢ = X ¡<br />

<br />

−1 ¢ <br />

<br />

(∇) = X ¡<br />

<br />

−1 ¢ <br />

(∇) <br />

<br />

<br />

(E.23)<br />

Aga<strong>in</strong> the summation cancels the superscript and subscript. The Kronecker delta symbol is written as<br />

X<br />

= <br />

(E.24)<br />

E.5 Generalized <strong>in</strong>ner product<br />

<br />

The generalized def<strong>in</strong>ition of an <strong>in</strong>ner product is<br />

= X <br />

<br />

(E.25)<br />

where is a unitary matrix called a covariant metric. The covariant metric transforms a contravariant to<br />

a covariant tensor. For example the matrix element of a covariant tensor can be written as<br />

= X <br />

(E.26)<br />

<br />

By association of the covariant metric with either of the vectors <strong>in</strong> the <strong>in</strong>ner product gives<br />

= X <br />

= X <br />

= X <br />

<br />

(E.27)<br />

Similarly it can be def<strong>in</strong>ed <strong>in</strong> terms of an orthogonal contravariant metric where<br />

= X <br />

<br />

(E.28)<br />

Then<br />

= X <br />

<br />

(E.29)<br />

Association of the contravariant metric with one of the vectors <strong>in</strong> the <strong>in</strong>ner product gives the <strong>in</strong>ner<br />

product<br />

= X = X = X <br />

(E.30)<br />

<br />

<br />

<br />

For most situations <strong>in</strong> this book the metric is diagonal and unitary.


528 APPENDIX E. TENSOR ALGEBRA<br />

E.6 Transformation properties of observables<br />

In physics, observables can be represented by spherical tensors which specify the angular momentum and<br />

parity characteristics of the observable, and the tensor rank is <strong>in</strong>dependent of the time dependence. The<br />

transformation properties of these tensors, coupled with their time-reversal <strong>in</strong>variance, specify the fundamental<br />

characteristics of the observables.<br />

Table 2 summarizes the transformation properties under rotation, spatial <strong>in</strong>version and time reversal<br />

for observables encountered <strong>in</strong> classical mechanics and electrodynamics. Note that observables can be scalar,<br />

vector, pseudovector, or second-order tensors, under rotation, and even or odd under either space <strong>in</strong>version<br />

or time <strong>in</strong>version. For example, <strong>in</strong> classical mechanics the <strong>in</strong>ertia tensor I relates the angular velocity vector<br />

ω to the angular momentum vector L by tak<strong>in</strong>g the <strong>in</strong>ner product L = I · ω. In general I is not diagonal and<br />

thus the angular momentum is not parallel to the angular velocity ω. A similar example <strong>in</strong> electrodynamics<br />

is the dielectric tensor K which relates the displacement field D to the electric field E by D = K · E. For<br />

anisotropic crystal media K is not diagonal lead<strong>in</strong>g to the electric field vectors E and D not be<strong>in</strong>g parallel.<br />

As discussed <strong>in</strong> chapter 7, Noether’s Theorem states that symmetries of the transformation properties lead<br />

to important conservation laws. The behavior of classical systems under rotation relates to the conservation<br />

of angular momentum, the behavior under spatial <strong>in</strong>version relates to parity conservation, and time-reversal<br />

<strong>in</strong>variance relates to conservation of energy. That is, conservative forces conserve energy and are time-reversal<br />

<strong>in</strong>variant.<br />

Table 2 :Transformation properties of scalar, vector, pseudovector, and tensor observables<br />

under rotation, spatial <strong>in</strong>version, and time reversal 2<br />

Physical Observable Rotation Space Time Name<br />

(Tensor rank) <strong>in</strong>version reversal<br />

1) <strong>Classical</strong> <strong>Mechanics</strong><br />

Mass density 0 Even Even Scalar<br />

K<strong>in</strong>etic energy 2 2 0 Even Even Scalar<br />

Potential energy () 0 Even Even Scalar<br />

Lagrangian 0 Even Even Scalar<br />

Hamiltonian 0 Even Even Scalar<br />

Gravitational potential 0 Even Even Scalar<br />

Coord<strong>in</strong>ate r 1 Odd Even Vector<br />

Velocity v 1 Odd Odd Vector<br />

Momentum p 1 Odd Odd Vector<br />

Angular momentum L = r × p 1 Even Odd Pseudovector<br />

Force F 1 Odd Even Vector<br />

Torque N = r × F 1 Even Even Pseudovector<br />

Gravitational field g 1 Odd Even Vector<br />

Inertia tensor I 2 Even Even Tensor<br />

Elasticity stress tensor T 2 Even Even Tensor<br />

2) Electromagnetism<br />

Charge density 0 Even Even Scalar<br />

Current density j 1 Odd Odd Vector<br />

Electric field E 1 Odd Even Vector<br />

Polarization P 1 Odd Even Vector<br />

Displacement D 1 Odd Even Vector<br />

Magnetic field B 1 Even Odd Pseudovector<br />

Magnetization M 1 Even Odd Pseudovector<br />

Magnetic field H 1 Even Odd Pseudovector<br />

Poynt<strong>in</strong>g vector S = E × H 1 Odd Odd Vector<br />

Dielectric tensor K 2 Even Even Tensor<br />

Maxwell stress tensor T 2 Even Even Tensor<br />

2 Basedontable6.1<strong>in</strong>"<strong>Classical</strong> Electrodynamics" 2 edition, by J.D. Jackson [Jac75]


Appendix F<br />

Aspects of multivariate calculus<br />

Multivariate calculus provides the framework for handl<strong>in</strong>g systems hav<strong>in</strong>g many variables associated with<br />

each of several bodies. It is assumed that the reader has studied l<strong>in</strong>ear differential equations plus multivariate<br />

calculus and thus has been exposed to the calculus used <strong>in</strong> classical mechanics. Chapter 5 of this book<br />

<strong>in</strong>troduced variational calculus which covers several important aspects of multivariate calculus such as Euler’s<br />

variational calculus and Lagrange multipliers. This appendix provides a brief review of a selection of other<br />

aspects of multivariate calculus that feature prom<strong>in</strong>ently <strong>in</strong> classical mechanics.<br />

F.1 Partial differentiation<br />

The extension of the derivative to multivariate calculus <strong>in</strong>volves use of partial derivatives. The partial<br />

derivative with respect to the variable of a multivariate function ( 1 2 ) <strong>in</strong>volves tak<strong>in</strong>g the<br />

normal one-variable derivative with respect to assum<strong>in</strong>g that the other − 1 variables are held constant.<br />

That is,<br />

∙ ¸<br />

( 1 2 ) (1 2 −1 ( + ) ) − ( 1 2 )<br />

= lim<br />

(F.1)<br />

<br />

→0<br />

<br />

where it will be assumed that the function () is a cont<strong>in</strong>uously-differentiable function to order, then<br />

all partial derivatives of that order or less are <strong>in</strong>dependent of the order <strong>in</strong> which they are performed. That<br />

is,<br />

2 ()<br />

= 2 ()<br />

(F.2)<br />

<br />

The cha<strong>in</strong> rule for partial differentiation gives that<br />

( 1 2 )<br />

<br />

=<br />

X<br />

=1<br />

() ()<br />

<br />

(F.3)<br />

The total differential of a multivariate function () is<br />

=<br />

X<br />

=1<br />

()<br />

<br />

(F.4)<br />

This can be extended to higher-order derivatives us<strong>in</strong>g the operator formalism<br />

µ<br />

<br />

<br />

<br />

() = 1 + + () = X ()<br />

1 <br />

1 1 <br />

F.2 L<strong>in</strong>ear operators<br />

The l<strong>in</strong>ear operator notation provides a powerful, elegant, and compact way to express, and apply, the<br />

equations of multivariate calculus; it is used extensively <strong>in</strong> mathematics and physics. The l<strong>in</strong>ear operators<br />

529<br />

(F.5)


530 APPENDIX F. ASPECTS OF MULTIVARIATE CALCULUS<br />

typically comprise partial derivatives that act on scalar, vector, or tensor fields. Table 1 lists a few<br />

elementary examples of the use of l<strong>in</strong>ear operators <strong>in</strong> this textbook. The first four l<strong>in</strong>ear operators <strong>in</strong>volve<br />

the widely used del operator ∇ to generate the gradient, divergence and curl as described <strong>in</strong> appendices <br />

and . The fifth and sixth l<strong>in</strong>ear operators act on the Lagrangian <strong>in</strong> Lagrangian mechanics applications.<br />

The f<strong>in</strong>al two l<strong>in</strong>ear operators act on the wavefunction for wave mechanics.<br />

Name Partial derivative Field Action<br />

Gradient<br />

∇ ≡ ˆ<br />

³<br />

<br />

+ ˆ <br />

+ ˆk ´<br />

Scalar potential E = ∇<br />

Divergence<br />

∇· ≡ ˆ <br />

+ ˆ <br />

+ ˆk <br />

³<br />

·<br />

´<br />

Vector field E ∇·E<br />

Curl<br />

∇× ≡ ˆ <br />

+ ˆ <br />

+ ˆk <br />

× Vector field E ∇ × E<br />

Laplacian<br />

∇ 2 = ∇·∇ ≡ 2<br />

<br />

+ 2<br />

2 <br />

+ 2<br />

2 2 Scalar potential ∇ 2 <br />

Euler-Lagrange Λ ≡ <br />

˙ <br />

− <br />

<br />

Scalar Lagrangian Λ =0<br />

Canonical momentum ≡ <br />

˙ <br />

Scalar Lagrangian ≡ <br />

Canonical momentum ≡ <br />

˙ <br />

Wavefunction Ψ Ψ ≡ <br />

Hamiltonian = ~ <br />

Wavefunction Ψ Ψ = ~ Ψ<br />

<br />

Table 1 examples of l<strong>in</strong>ear operators used <strong>in</strong> this textbook.<br />

˙ <br />

Ψ<br />

˙ <br />

= Ψ<br />

There are three ways of express<strong>in</strong>g operations such as addition, multiplication, transposition or <strong>in</strong>version<br />

of operations that are completely equivalent because they all are based on the same pr<strong>in</strong>ciples of l<strong>in</strong>ear<br />

algebra. For example, a transformation O act<strong>in</strong>g on a vector A can produced the vector B. The simplest<br />

way to express this transformation is <strong>in</strong> terms of components<br />

3X<br />

= <br />

=1<br />

(F.6)<br />

Another way is to use matrix mechanics where the 3 × 3 matrix (O) transforms the column vector (A) to<br />

the column vector (B), thatis,<br />

(B)=(O)(A)<br />

(F.7)<br />

The third approach is to assume an operator O acts on the vector A<br />

B = OA<br />

(F.8)<br />

In classical mechanics, and quantum mechanics, these three equivalent approaches are used and exploited<br />

extensively and <strong>in</strong>terchangeably. In particular the rules of matrix manipulation, that are given <strong>in</strong> appendix<br />

are synonymous, and equivalent to, those that apply for operator manipulation. If the operator is complex<br />

then the operator properties are summarized as follows.<br />

The generalization of the transpose for complex operators is the Hermitian conjugate †<br />

† = ∗ <br />

(F.9)<br />

Note also that<br />

O † =( ∗ ) =( ) ∗<br />

(F.10)<br />

The generalization of a symmetric matrix is Hermitian, that is, is equal to its Hermitian conjugate<br />

† = ∗ = <br />

(F.11)<br />

For a real matrix the complex conjugation has no effect so the matrix is real and symmetric.<br />

The generalization of orthogonal is unitary for which the operator is unitary if it is non-s<strong>in</strong>gular and<br />

which implies<br />

−1 = †<br />

† = = † <br />

(F.12)<br />

(F.13)


F.3. TRANSFORMATION JACOBIAN 531<br />

F.3 Transformation Jacobian<br />

The Jacobian determ<strong>in</strong>ant, which is usually called the Jacobian, is used extensively <strong>in</strong> mechanics for both<br />

rotational and translational coord<strong>in</strong>ate transformations. The Jacobian determ<strong>in</strong>ant is def<strong>in</strong>ed as be<strong>in</strong>g the<br />

ratio of the -dimensional volume element 1 2 <strong>in</strong>onecoord<strong>in</strong>atesystem,tothevolumeelement<br />

1 2 <strong>in</strong> the second coord<strong>in</strong>ate system. That is<br />

¯ 1<br />

<br />

1<br />

F.3.1<br />

( 1 2 ) ≡ 1 2 <br />

1 2 <br />

=<br />

Transformation of <strong>in</strong>tegrals:<br />

¯<br />

1<br />

1 2<br />

2 2<br />

1 2<br />

.<br />

<br />

1<br />

.<br />

<br />

2<br />

<br />

.<br />

<br />

<br />

2<br />

<br />

.<br />

<br />

<br />

¯¯¯¯¯¯¯¯¯¯<br />

(F.14)<br />

Consider a coord<strong>in</strong>ate transformation for the <strong>in</strong>tegral of the function ( 1 2 ) to the <strong>in</strong>tegral of a<br />

function ( 1 2 ) where = ( 1 2 ) The coord<strong>in</strong>ate transformation of the <strong>in</strong>tegral equation<br />

can be expressed <strong>in</strong> terms of the Jacobian ( 1 2 )<br />

Z<br />

Z<br />

( 1 2 ) 1 2 = ( 1 2 ) 1 2 =<br />

(F.15)<br />

F.3.2<br />

Z<br />

( 1 2 ) 1 2 <br />

1 2 <br />

1 2 =<br />

Transformation of differential equations:<br />

Z<br />

( 1 2 )( 1 2 ) 1 2 <br />

The differential cross sections for scatter<strong>in</strong>g can be def<strong>in</strong>ed either by the number of a def<strong>in</strong>ite k<strong>in</strong>d of<br />

particle/per event, go<strong>in</strong>g <strong>in</strong>to the volume element <strong>in</strong> momentum space 1 2 3 or by the number go<strong>in</strong>g<br />

<strong>in</strong>to the solid angle element hav<strong>in</strong>g momentum between and + . That is, the first def<strong>in</strong>ition can be<br />

written as a differential equation<br />

3 ( 1 2 3 )<br />

1 2 3 = 3 ( 1 () 2 () 3 ()) ( 1 2 3 )<br />

1 2 3<br />

1 2 3 ( ) <br />

(F.16)<br />

As shown <strong>in</strong> table 4, 1 2 3 = 2 s<strong>in</strong> that is, the Jacobian equals 2 s<strong>in</strong> Thus equation 16<br />

can be written as<br />

3 ∙<br />

( 1 2 3 )<br />

3 2¸<br />

1 2 3 =<br />

(s<strong>in</strong> ) = 2 ( )<br />

Ω (F.17)<br />

1 2 3 1 2 3 Ω<br />

The differential cross section is def<strong>in</strong>ed by<br />

2 ( )<br />

Ω<br />

≡<br />

3 <br />

1 2 3<br />

2<br />

(F.18)<br />

where the 2 factor is absorbed <strong>in</strong>to the cross section and the solid angle term is factored out<br />

F.3.3<br />

Properties of the Jacobian:<br />

In classical mechanics the Jacobian often is extended from 3 dimensions to -dimensional transformations.<br />

The Jacobian is unity for unitary transformations such as rotations and l<strong>in</strong>ear translations which implies that<br />

the volume element is preserved. It will be shown that this also is true for a certa<strong>in</strong> class of transformations<br />

<strong>in</strong> classical mechanics that are called canonical transformations. The Jacobian transforms the local density<br />

to be correct for any scale transformations such as transform<strong>in</strong>g l<strong>in</strong>ear dimensions from centimeters to <strong>in</strong>ches.<br />

F.1 Example: Jacobian for transform from cartesian to spherical coord<strong>in</strong>ates<br />

Consider the transform <strong>in</strong> the three-dimensional <strong>in</strong>tegral R ( 1 2 3 ) 1 2 3 under transformation<br />

from cartesian coord<strong>in</strong>ates ( 1 2 3 ) to spherical coord<strong>in</strong>ates ( ) The transformation is governed by


532 APPENDIX F. ASPECTS OF MULTIVARIATE CALCULUS<br />

the geometric relations 1 = s<strong>in</strong> cos , 2 = s<strong>in</strong> s<strong>in</strong> , 3 = cos . For this transformation the Jacobian<br />

determ<strong>in</strong>ant equals<br />

s<strong>in</strong> cos cos cos − s<strong>in</strong> s<strong>in</strong> <br />

( ) =<br />

s<strong>in</strong> s<strong>in</strong> cos s<strong>in</strong> s<strong>in</strong> cos <br />

¯ cos − s<strong>in</strong> 0 ¯ = 2 s<strong>in</strong> <br />

Thus the three-dimensional volume <strong>in</strong>tegral transforms to<br />

Z<br />

Z<br />

Z<br />

( 1 2 3 ) 1 2 3 = ( )( ) = ( ) 2 s<strong>in</strong> <br />

which is the well-known volume <strong>in</strong>tegral <strong>in</strong> spherical coord<strong>in</strong>ates.<br />

F.4 Legendre transformation<br />

Hamiltonian mechanics can be derived directly from Lagrange mechanics by consider<strong>in</strong>g the Legendre transformation<br />

between the conjugate variables (q ˙q) and (q p) Such a derivation is of considerable importance<br />

<strong>in</strong> that it shows that Hamiltonian mechanics is based on the same variational pr<strong>in</strong>ciples as those<br />

used to derive Lagrangian mechanics; that is d’Alembert’s Pr<strong>in</strong>ciple or Hamilton’s Pr<strong>in</strong>ciple. The general<br />

problem of convert<strong>in</strong>g Lagrange’s equations <strong>in</strong>to the Hamiltonian form h<strong>in</strong>ges on the <strong>in</strong>version of equation<br />

(83) that def<strong>in</strong>es the generalized momentum p This <strong>in</strong>version is simplified by the fact that (83) is the first<br />

partial derivative of the Lagrangian (q ˙q t) which is a scalar function.<br />

Consider transformations between two functions (u w) and (v w) where u and v are the active<br />

variables related by the functional form<br />

v = ∇ u (u w)<br />

(F.19)<br />

and where w designates passive variables and ∇ u (u w) is the first-order derivative of (u w) , i.e. the<br />

gradient, with respect to the components of the vector u. The Legendre transform states that the <strong>in</strong>verse<br />

formula can always be written <strong>in</strong> the form<br />

u = ∇ v (v w)<br />

where the function (v w) is related to (u w) by the symmetric relation<br />

(v w)+F(u w) =u · v<br />

and where the scalar product u · v = P <br />

=1 .<br />

Furthermore the derivatives with respect to all the passive variables { } are related by<br />

∇ w (u w) =−∇ w<br />

(v w)<br />

(F.20)<br />

(F.21)<br />

(F.22)<br />

The relationship between the functions (u w) and (v w) is symmetrical and each is said to be the<br />

Legendre transform of the other.<br />

Problems<br />

1. Below you will f<strong>in</strong>d a set of <strong>in</strong>tegrals. Your teach<strong>in</strong>g assistant will divide you <strong>in</strong>to groups and each group will<br />

be assigned one <strong>in</strong>tegral to work on. Once your group has solved the <strong>in</strong>tegral, write the solution on the board<br />

<strong>in</strong> the space provided by the teach<strong>in</strong>g assistant.<br />

(a) R 2<br />

0<br />

R 4<br />

0<br />

R cos <br />

0<br />

2 s<strong>in</strong> <br />

(b) R ¡ ṙ<br />

− ¢ r ˙<br />

2<br />

(c) R A · a where A = ˆ + ˆ + ˆk and is the sphere 2 + 2 + 2 =9.<br />

(d) R (∇ × A) · a where A = ˆ + ˆ + ˆk and is the surface def<strong>in</strong>ed by the paraboloid =1− 2 − 2 ,<br />

where ≥ 0.


Appendix G<br />

Vector differential calculus<br />

This appendix reviews vector differential calculus which is used extensively <strong>in</strong> both classical mechanics and<br />

electromagnetism.<br />

G.1 Scalar differential operators<br />

G.1.1<br />

Scalar field<br />

Differential operators like time ¡ <br />

¢<br />

do not change the rotational properties of scalars or proper vectors. A<br />

scalar operator act<strong>in</strong>g on a scalar field (), <strong>in</strong> a rotated coord<strong>in</strong>ated frame 0 ( 0 0 0 ) is unchanged.<br />

0<br />

= <br />

<br />

(G.1)<br />

G.1.2<br />

Vector field<br />

Similarly for a proper vector field<br />

0 <br />

= X <br />

<br />

<br />

<br />

(G.2)<br />

That is, differentiation of scalar or vector fields with respect to a scalar operator does not change the<br />

rotational behavior. In particular, the scalar differentials of vectors cont<strong>in</strong>ue to obey the rules of ord<strong>in</strong>ary<br />

proper vectors. The scalar operator <br />

is used for calculation of velocity or acceleration.<br />

G.2 Vector differential operators <strong>in</strong> cartesian coord<strong>in</strong>ates<br />

Vector differential operators, such as the gradient operator, are important <strong>in</strong> physics. The action of vector<br />

operators differ along different orthogonal axes.<br />

G.2.1<br />

Scalar field<br />

Consider a cont<strong>in</strong>uous, s<strong>in</strong>gle-valued scalar function ( ) S<strong>in</strong>ce<br />

0 = <br />

(G.3)<br />

then the partial differential with respect to one component of the vector x 0 gives<br />

The <strong>in</strong>verse rotation gives that<br />

0<br />

0 = X<br />

<br />

= X <br />

<br />

<br />

<br />

0 <br />

0 <br />

(G.4)<br />

(G.5)<br />

533


534 APPENDIX G. VECTOR DIFFERENTIAL CALCULUS<br />

Therefore<br />

<br />

0 <br />

= X <br />

<br />

0 <br />

0 <br />

= X <br />

= <br />

(G.6)<br />

Thus<br />

0<br />

0 = X<br />

<br />

<br />

<br />

<br />

(G.7)<br />

That is the vector derivative act<strong>in</strong>g of a scalar field transforms like a proper vector.<br />

Def<strong>in</strong>e the gradient, or ∇ operator, as<br />

∇ ≡ X <br />

be <br />

<br />

<br />

(G.8)<br />

where be is the unit vector along the axis. In cartesian coord<strong>in</strong>ates, the del vector operator is,<br />

∇ ≡ b i <br />

+b j + b k <br />

(G.9)<br />

The gradient was applied to the gravitational and electrostatic potential to derive the correspond<strong>in</strong>g field.<br />

For example, for electrostatics it was shown that the gradient of the scalar electrostatic potential field can<br />

be written <strong>in</strong> cartesian coord<strong>in</strong>ates as<br />

E = −∇<br />

(G.10)<br />

Note that the gradient of a scalar field produces a vector field. You are familiar with this if you are a skier<br />

<strong>in</strong> that the gravitational force pulls you down the l<strong>in</strong>e of steepest descent for the ski slope.<br />

G.2.2<br />

Vector field<br />

Another possible operation for the del operator is the scalar product with a vector. Us<strong>in</strong>g the def<strong>in</strong>ition of<br />

a scalar product <strong>in</strong> cartesian coord<strong>in</strong>ates gives<br />

∇ · A = b i ·bi <br />

+b j ·bj <br />

+ k b · bk <br />

<br />

= <br />

+ <br />

+ <br />

<br />

(G.11)<br />

This scalar derivative of a vector field is called the divergence. Note that the scalar product produces a<br />

scalar field which is <strong>in</strong>variant to rotation of the coord<strong>in</strong>ate axes.<br />

The vector product of the del operator with another vector, is called the curl which is used extensively<br />

<strong>in</strong> physics. It can be written <strong>in</strong> the determ<strong>in</strong>ant form<br />

b i b j k b<br />

∇ × A =<br />

<br />

(G.12)<br />

<br />

¯ <br />

¯¯¯¯¯¯ <br />

By contrast to the scalar product, both the gradient of a scalar field, and the vector product, are vector<br />

fields for which the components along the coord<strong>in</strong>ate axes transform <strong>in</strong> a specific manner, such as to keep the<br />

length of the vector constant, as the coord<strong>in</strong>ate frame is rotated. The gradient, scalar and vector products<br />

with the ∇ operator are the first order derivatives of fields that occur most frequently <strong>in</strong> physics.<br />

<strong>Second</strong> derivatives of fields also are used. Let us consider some possible comb<strong>in</strong>ations of the product of<br />

two del operators.<br />

1) ∇· (∇ )=∇ 2 <br />

The scalar product of two del operators is a scalar under rotation. Evaluat<strong>in</strong>g the scalar product <strong>in</strong><br />

cartesian coord<strong>in</strong>ates gives<br />

µ<br />

b<br />

i<br />

+b j + k b µ<br />

· b<br />

i<br />

+b j <br />

+ k b <br />

= 2 <br />

2 + 2 <br />

2<br />

This also can be obta<strong>in</strong>ed without confusion by writ<strong>in</strong>g this product as;<br />

+ 2 <br />

2<br />

(G.13)<br />

∇· (∇ )=∇ · ∇ =(∇ · ∇) (G.14)


G.3. VECTOR DIFFERENTIAL OPERATORS IN CURVILINEAR COORDINATES 535<br />

where the scalar product of the del operator is a scalar, called the Laplacian ∇ 2 given by<br />

∇ · ∇ = ∇ 2 ≡ 2<br />

2 + 2<br />

2 + 2<br />

2<br />

(G.15)<br />

The Laplacian operator is encountered frequently <strong>in</strong> physics.<br />

2) ∇× (∇ )=0<br />

Note that the vector product of two identical vectors<br />

A × A =0<br />

(G.16)<br />

Therefore<br />

∇× (∇ )=0 (G.17)<br />

This can be confirmed by evaluat<strong>in</strong>g the separate components along each axis.<br />

3) ∇· (∇ × A) =0<br />

This is zero because the cross-product is perpendicular to ∇ × A and thus the dot product is zero.<br />

4) ∇× (∇ × A) =∇· (∇ · A) −∇ 2 A<br />

The identity<br />

A × (B × C) =B (A · C) − (A · B) C<br />

(G.18)<br />

canbeusedtogive<br />

∇× (∇ × A) =∇· (∇ · A) −∇ 2 A (G.19)<br />

s<strong>in</strong>ce ∇ · ∇ = ∇ 2 <br />

There are pitfalls <strong>in</strong> the discussion of second derivatives <strong>in</strong> that it is assumed that both del operators<br />

operate on the same variable, otherwise the results are different.<br />

G.3 Vector differential operators <strong>in</strong> curvil<strong>in</strong>ear coord<strong>in</strong>ates<br />

As discussed <strong>in</strong> Appendix there are many situations where the symmetries make it more convenient to use<br />

orthogonal curvil<strong>in</strong>ear coord<strong>in</strong>ate systems rather than cartesian coord<strong>in</strong>ates. Thus it is necessary to extend<br />

vector derivatives from cartesian to curvil<strong>in</strong>ear coord<strong>in</strong>ates. Table 1 can be used for express<strong>in</strong>g vector<br />

derivatives <strong>in</strong> curvil<strong>in</strong>ear coord<strong>in</strong>ate systems.<br />

G.3.1<br />

Gradient:<br />

The gradient <strong>in</strong> curvil<strong>in</strong>ear coord<strong>in</strong>ates is<br />

∇ = 1 1<br />

<br />

1<br />

ˆq 1 + 1 2<br />

<br />

2<br />

ˆq 2 + 1 3<br />

<br />

3<br />

ˆq 3<br />

(G.20)<br />

where the coefficients are listed <strong>in</strong> table 1.<br />

For cyl<strong>in</strong>drical coord<strong>in</strong>ates this becomes<br />

In spherical coord<strong>in</strong>ates<br />

∇ = <br />

ˆρ + 1 <br />

<br />

ˆϕ +<br />

<br />

ẑ<br />

∇ = <br />

ˆr + 1 <br />

ˆθ + 1 <br />

s<strong>in</strong> ˆϕ<br />

(G.21)<br />

(G.22)


536 APPENDIX G. VECTOR DIFFERENTIAL CALCULUS<br />

G.3.2<br />

Divergence:<br />

Thedivergencecanbeexpressedas<br />

∇ · A = 1<br />

1 2 3<br />

∙ <br />

1<br />

( 1 2 3 )+<br />

<br />

2<br />

( 2 3 1 )+<br />

¸<br />

( 3 1 2 )<br />

3<br />

(G.23)<br />

In cyl<strong>in</strong>drical coord<strong>in</strong>ates the divergence is<br />

∇ · A = 1 <br />

( )+ 1 <br />

<br />

+ <br />

<br />

= <br />

+ <br />

+ 1 <br />

+ <br />

<br />

In spherical coord<strong>in</strong>ates the divergence is<br />

∇ · A = 1 ∙ ¡<br />

<br />

2 2 s<strong>in</strong> ¢ + s<strong>in</strong> <br />

( s<strong>in</strong> )+ ¸<br />

( )<br />

(G.24)<br />

(G.25)<br />

G.3.3<br />

Curl:<br />

∇ × A = 1<br />

1 2 3<br />

¯¯¯¯¯¯<br />

1ˆq 1 2ˆq 2 3ˆq 3<br />

<br />

1<br />

<br />

2<br />

<br />

3<br />

1 1 2 2 3 3<br />

¯¯¯¯¯¯<br />

In cyl<strong>in</strong>drical coord<strong>in</strong>ates the curl is<br />

∇ × A = 1 ˆρ ˆϕ ẑ<br />

<br />

<br />

<br />

¯ <br />

¯¯¯¯¯¯<br />

(G.26)<br />

(G.27)<br />

In spherical coord<strong>in</strong>ates the curl is<br />

∇ × A = 1<br />

2 s<strong>in</strong> <br />

¯<br />

ˆr ˆθ s<strong>in</strong> ˆϕ<br />

<br />

<br />

<br />

<br />

<br />

<br />

s<strong>in</strong> <br />

¯¯¯¯¯¯<br />

(G.28)<br />

G.3.4<br />

Laplacian:<br />

Tak<strong>in</strong>g the divergence of the gradient of a scalar gives<br />

∙ µ <br />

∇ 2 1 2 3 <br />

= ∇ · ∇ =<br />

+ µ <br />

3 1 <br />

+ µ ¸<br />

1 2 <br />

1 2 3 1 1 1 2 2 2 3 3 3<br />

The Laplacian of a scalar function <strong>in</strong> cyl<strong>in</strong>drical coord<strong>in</strong>ates is<br />

∇ 2 = 1 <br />

<br />

µ<br />

<br />

<br />

<br />

+ 1 2 2 <br />

2 + 2 <br />

2<br />

The Laplacian of a scalar function <strong>in</strong> spherical coord<strong>in</strong>ates is<br />

∇ 2 = 1 µ<br />

<br />

2 2 <br />

µ<br />

1 <br />

+<br />

2 s<strong>in</strong> <br />

s<strong>in</strong> <br />

<br />

+<br />

1 2 <br />

2 s<strong>in</strong> 2<br />

(G.29)<br />

(G.30)<br />

(G.31)<br />

The gradient, divergence, curl and Laplacian are used extensively <strong>in</strong> curvil<strong>in</strong>ear coord<strong>in</strong>ate systems when<br />

deal<strong>in</strong>g with vector fields <strong>in</strong> Newtonian mechanics, electromagnetism, and fluid flow.


Appendix H<br />

Vector <strong>in</strong>tegral calculus<br />

Field equations, such as for electromagnetic and gravitational fields, require both l<strong>in</strong>e <strong>in</strong>tegrals, and surface<br />

<strong>in</strong>tegrals, of vector fields to evaluate potential, flux and circulation. These require use of the gradient, the<br />

Divergence Theorem and Stokes Theorem which are discussed <strong>in</strong> the follow<strong>in</strong>g sections.<br />

H.1 L<strong>in</strong>e <strong>in</strong>tegral of the gradient of a scalar field<br />

The change ∆ <strong>in</strong> a scalar field for an <strong>in</strong>f<strong>in</strong>itessimal step l along a path can be written as<br />

∆ =(∇ ) · l<br />

(H.1)<br />

s<strong>in</strong>ce the gradient of that is, ∇ istherateofchangeof with l Discussions of gravitational and<br />

electrostatic potential show that the l<strong>in</strong>e <strong>in</strong>tegral between po<strong>in</strong>ts and is given <strong>in</strong> terms of the del operator<br />

by<br />

Z <br />

− = (∇ ) · l<br />

(H.2)<br />

This relates the difference <strong>in</strong> values of a scalar field at two po<strong>in</strong>ts to the l<strong>in</strong>e <strong>in</strong>tegral of the dot product of<br />

the gradient with the element of the l<strong>in</strong>e <strong>in</strong>tegral.<br />

H.2 Divergence theorem<br />

<br />

H.2.1<br />

Flux of a vector field for Gaussian surface<br />

Consider the flux Φ of a vector field F for a closed surface, usually<br />

called a Gaussian surface, shown <strong>in</strong> figure 1.<br />

I<br />

Φ = F · S<br />

(H.3)<br />

<br />

If the enclosed volume is cut <strong>in</strong> to two pieces enclosed by surfaces<br />

1 = + and 2 = + . The flux through the surface <br />

common to both 1 and 2 are equal and <strong>in</strong> the same direction. Then<br />

the net flux through the sum of 1 and 2 is given by<br />

I I<br />

I<br />

F · S + F · S = F · S<br />

(H.4)<br />

1 2 <br />

s<strong>in</strong>ce the contributions of the common surface cancel <strong>in</strong> that the<br />

flux out of 1 is equal and opposite to the flux <strong>in</strong>to 2 over the surface<br />

That is, <strong>in</strong>dependent of how many times the volume enclosed by<br />

is subdivided, the net flux for the sum of all the Gaussian surfaces<br />

enclos<strong>in</strong>g these subdivisions of the volume, still equals H F · S<br />

537<br />

S a<br />

S 2 S 1<br />

cut<br />

S b<br />

Sab<br />

F<br />

V 1 V 2<br />

Figure H.1: A volume V enclosed<br />

byaclosedsurfaceSiscut<strong>in</strong>totwo<br />

pieces at the surface S This gives<br />

V 1 enclosed by S 1 and V 1 enclosed<br />

by S 2 <br />

F


538 APPENDIX H. VECTOR INTEGRAL CALCULUS<br />

Consider that the volume enclosed by is subdivided <strong>in</strong>to subdivisions where →∞ then even<br />

though H <br />

F · S → 0 as →∞, the sum over surfaces of all the <strong>in</strong>f<strong>in</strong>itessimal volumes rema<strong>in</strong>s unchanged<br />

I<br />

Φ =<br />

<br />

F · S =<br />

→∞<br />

X<br />

<br />

I<br />

<br />

F · S<br />

(H.5)<br />

Thus we can take the limit of a sum of an <strong>in</strong>f<strong>in</strong>ite number of <strong>in</strong>f<strong>in</strong>itessimal volumes as is needed to obta<strong>in</strong> a<br />

differential form. The surface <strong>in</strong>tegral for each <strong>in</strong>f<strong>in</strong>itessimal volume will equal zero which is not useful, that<br />

is H <br />

F · S → 0 as →∞ However, the flux per unit volume has a f<strong>in</strong>ite value as →∞ This ratio is<br />

called the divergence of the vector field;<br />

H<br />

<br />

F = <br />

F · S<br />

∆ →0<br />

(H.6)<br />

∆ <br />

where ∆ is the <strong>in</strong>f<strong>in</strong>itessimal volume enclosed by surface Thedivergenceofthevectorfield is a scalar<br />

quantity.<br />

Thus the sum of flux over all <strong>in</strong>f<strong>in</strong>itessimal subdivisions of the volume enclosed by a closed surface <br />

equals<br />

I<br />

H<br />

→∞<br />

X<br />

→∞<br />

<br />

Φ = F · S =<br />

<br />

F · S X<br />

∆ = F∆ <br />

(H.7)<br />

∆ <br />

In the limit →∞ ∆ → 0 this becomes the <strong>in</strong>tegral;<br />

I Z<br />

Φ = F · S =<br />

<br />

<br />

<br />

<br />

<br />

<br />

F<br />

This is called the Divergence Theorem or Gauss’s Theorem. To avoid confusion with Gauss’s law <strong>in</strong> electrostatics,<br />

it will be referred to as the Divergence theorem.<br />

(H.8)<br />

H.2.2<br />

Divergence <strong>in</strong> cartesian coord<strong>in</strong>ates.<br />

Consider the special case of an <strong>in</strong>f<strong>in</strong>itessimal rectangular box, size<br />

∆ ∆ ∆ shown <strong>in</strong> figure 2 Consider the net flux for the component<br />

enter<strong>in</strong>g the surface ∆∆ at location ( ).<br />

µ<br />

∆Φ <br />

=<br />

+ ∆<br />

2<br />

<br />

+ ∆<br />

2<br />

<br />

<br />

∆∆<br />

<br />

(H.9)<br />

z<br />

x,y,z<br />

F z<br />

The net flux of the z component out of the surface at + ∆ is<br />

µ<br />

∆Φ <br />

= + ∆ <br />

<br />

∆∆<br />

+ ∆<br />

2<br />

<br />

+ ∆<br />

2<br />

<br />

<br />

(H.10)<br />

y<br />

Thus the net flux out of the box due to the z component of F is<br />

∆Φ = ∆Φ <br />

<br />

− ∆Φ <br />

<br />

= <br />

∆∆∆<br />

Add<strong>in</strong>g the similar and components for ∆Φ gives<br />

µ <br />

∆Φ =<br />

+ <br />

+ <br />

<br />

∆∆∆<br />

<br />

(H.11)<br />

(H.12)<br />

Figure H.2: Computation of flux<br />

out of an <strong>in</strong>f<strong>in</strong>itessimal rectangular<br />

box, ∆ ∆ ∆<br />

x<br />

This gives that the divergence of the vector field F is<br />

H<br />

F = ∆ →0<br />

<br />

F · S<br />

∆ <br />

=<br />

µ <br />

+ <br />

+ <br />

<br />

<br />

(H.13)


H.2. DIVERGENCE THEOREM 539<br />

s<strong>in</strong>ce ∆ = ∆∆∆ But the right hand side of the equation equals the scalar product ∇ · F that is,<br />

F = ∇ · F<br />

(H.14)<br />

The divergence is a scalar quantity. The physical mean<strong>in</strong>g of the divergence is that it gives the net flux per<br />

unit volume flow<strong>in</strong>g out of an <strong>in</strong>f<strong>in</strong>itessimal volume. A positive divergence corresponds to a net outflow of<br />

flux from the <strong>in</strong>f<strong>in</strong>itessimal volume at any location while a negative divergence implies a net <strong>in</strong>flow of flux<br />

to this <strong>in</strong>f<strong>in</strong>itessimal volume.<br />

It was shown that for an <strong>in</strong>f<strong>in</strong>itessimal rectangular box<br />

∆Φ =<br />

µ <br />

+ <br />

+ <br />

<br />

<br />

∆∆∆ = ∇ · F∆<br />

Integrat<strong>in</strong>g over the f<strong>in</strong>ite volume enclosed by the surface gives<br />

I Z<br />

Φ = F · S = ∇ · F<br />

This is another way of express<strong>in</strong>g the Divergence theorem<br />

I Z<br />

Φ = F · S =<br />

<br />

<br />

<br />

<br />

<br />

<br />

F<br />

(H.15)<br />

(H.16)<br />

(H.17)<br />

The divergence theorem, developed by Gauss, is of considerable importance, it relates the surface <strong>in</strong>tegral of<br />

a vector field, that is, the outgo<strong>in</strong>g flux, to a volume <strong>in</strong>tegral of ∇ · F over the enclosed volume.<br />

H.1 Example: Maxwell’s Flux Equations<br />

As an example of the usefulness of this relation, consider the Gauss’s law for the flux <strong>in</strong> Maxwell’s<br />

equations.<br />

Gauss’ Law for the electric field<br />

I<br />

Φ =<br />

<br />

<br />

But the divergence relation gives that<br />

I<br />

Φ = E · S =<br />

Comb<strong>in</strong><strong>in</strong>g these gives<br />

I<br />

<br />

<br />

<br />

Z<br />

E · S =<br />

E · dS = 1 0<br />

Z<br />

<br />

<br />

Z<br />

<br />

<br />

<br />

<br />

d<br />

∇ · E<br />

∇ · E = 1 0<br />

Z<br />

<br />

<br />

This is true <strong>in</strong>dependent of the shape of the surface or enclosed volume, lead<strong>in</strong>g to the differential form<br />

of Maxwell’s first law, that is Gauss’s law for the electric field.<br />

<br />

∇ · E = 0<br />

The differential form of Gauss’s law relates ∇ · E tothechargedensity at that same location. This is<br />

much easier to evaluate than a surface and volume <strong>in</strong>tegral required us<strong>in</strong>g the <strong>in</strong>tegral form of Gauss’s law.<br />

Gauss’s law for magnetism<br />

I<br />

Us<strong>in</strong>g the divergence theorem gives that<br />

I<br />

Φ =<br />

Φ =<br />

<br />

<br />

<br />

<br />

Z<br />

B · S =<br />

B · S =0<br />

<br />

<br />

∇ · B =0


540 APPENDIX H. VECTOR INTEGRAL CALCULUS<br />

This is true <strong>in</strong>dependent of the shape of the Gaussian surface lead<strong>in</strong>g to the differential form of Gauss’s law<br />

for B<br />

∇ · B =0<br />

That is, the local value of the divergence of B is zero everywhere.<br />

H.2 Example: Buoyancy forces <strong>in</strong> fluids<br />

Buoyancy <strong>in</strong> fluids provides an example of the use of flux <strong>in</strong> physics. Consider a fluid of density ()<br />

<strong>in</strong> a gravitational field ¯() =−()ˆ where the axis po<strong>in</strong>ts <strong>in</strong> the opposite direction to the gravitational<br />

force. Pressure equals force per unit area and is a scalar quantity. For a conservative fluid system, <strong>in</strong> static<br />

equilibrium, the net work done per unit area for an <strong>in</strong>f<strong>in</strong>itessimal displacement is zero. The net pressure<br />

force per unit area is the difference ( +)− () =∇ · while the net change <strong>in</strong> gravitational potential<br />

energy is ()¯() · . Thus energy conservation gives<br />

which can be expanded as<br />

[∇ + ()ḡ(z)] · r =0<br />

<br />

= −()() ()<br />

<br />

<br />

= <br />

=0<br />

Integrat<strong>in</strong>g the net forces normal to the surface over any closed surface enclos<strong>in</strong>g an empty volume, <strong>in</strong>side<br />

the fluid, gives a net buoyancy force on this volume that simplifies us<strong>in</strong>g the Divergence theorem<br />

I I<br />

I Z µ <br />

F · S= Ŝ · S = =<br />

+ <br />

+ <br />

<br />

<br />

Us<strong>in</strong>g equations leads to the net buoyancy force<br />

I Z<br />

<br />

F · S=<br />

<br />

<br />

<br />

<br />

= − Z<br />

<br />

()()<br />

The right hand side of this equation equals m<strong>in</strong>us the weight of the displaced fluid. That is, the buoyancy force<br />

equals the weight of the fluid displaced by the empty volume. Note that this proof applies both to compressible<br />

fluids, where the density depends on pressure, as well as to <strong>in</strong>compressible fluids where the density is constant.<br />

It also applies to situations where local gravity is position dependent. If an object of mass is completely<br />

submerged then the net force on the object is − R ()() If the object floats on the surface<br />

<br />

of a fluid then the buoyancy force must be calculated separately for the volume under the fluid surface and<br />

the upper volume above the fluid surface. The buoyancy due to displaced air usually is negligible s<strong>in</strong>ce the<br />

density of air is about 10 −3 times that of fluids such as water.<br />

H.3 Stokes Theorem<br />

H.3.1<br />

The curl<br />

Maxwell’s laws relate the circulation of the field around a closed loop to the rate of change of flux through<br />

the surface bounded by the closed loop. It is possible to write these <strong>in</strong>tegral equations <strong>in</strong> a differential form<br />

as follows.<br />

Consider the l<strong>in</strong>e <strong>in</strong>tegral around a closed loop shown <strong>in</strong> figure 3.<br />

If this area is subdivided <strong>in</strong>to two areas enclosed by loops 1 and 2 , then the sum of the l<strong>in</strong>e <strong>in</strong>tegrals<br />

is the same<br />

I I I<br />

F · l = F · l + F · l<br />

(H.18)<br />

<br />

1 2<br />

because the contributions along the common boundary cancel s<strong>in</strong>ce they are taken <strong>in</strong> opposite directions if<br />

1 and 2 both are taken <strong>in</strong> the same direction. Note that the l<strong>in</strong>e <strong>in</strong>tegral, and correspond<strong>in</strong>g enclosed area,


H.3. STOKES THEOREM 541<br />

are vector quantities related by the right-hand rule and this must be taken <strong>in</strong>to account when subdivid<strong>in</strong>g<br />

the area. Thus the area can be subdivided <strong>in</strong>to an <strong>in</strong>f<strong>in</strong>ite number of pieces for which<br />

I<br />

→∞<br />

X<br />

I<br />

H<br />

→∞<br />

X<br />

<br />

F · l = F · l =<br />

<br />

F · l<br />

<br />

<br />

∆S · bn ∆S · bn<br />

(H.19)<br />

<br />

where ∆S is the <strong>in</strong>f<strong>in</strong>itessimal area bounded by the closed sub-loop and ∆S · bn is the normal component<br />

of this area po<strong>in</strong>t<strong>in</strong>g along the bn direction which is the direction along which the l<strong>in</strong>e <strong>in</strong>tegral po<strong>in</strong>ts.<br />

The component of the curl of the vector function along the direction<br />

bn is def<strong>in</strong>ed to be<br />

H<br />

C<br />

→∞<br />

X<br />

<br />

(F) · bn ≡ <br />

F · l<br />

∆→0 (H.20)<br />

∆S · bn<br />

Thusthel<strong>in</strong>e<strong>in</strong>tegralcanbewrittenas<br />

I<br />

H<br />

→∞<br />

X<br />

<br />

F · l =<br />

<br />

F · l<br />

<br />

∆S<br />

· bn ∆S · bn<br />

Z<br />

= [(F) · bn] S · bn<br />

<br />

<br />

(H.21)<br />

The product bn · bn = 1, that is, this is true <strong>in</strong>dependent of the<br />

direction of the <strong>in</strong>f<strong>in</strong>itessimal loop. Thus the above relation leads<br />

to Stokes Theorem<br />

I Z<br />

<br />

F · l =<br />

<br />

<br />

<br />

<br />

(F) · S<br />

(H.22)<br />

This relates the l<strong>in</strong>e <strong>in</strong>tegral to a surface <strong>in</strong>tegral over a surface<br />

bounded by the loop.<br />

Figure H.3: The circulation around a<br />

path is equal to the sum of the circulations<br />

around subareas made by subdivid<strong>in</strong>g<br />

the area.<br />

H.3.2<br />

Curl <strong>in</strong> cartesian coord<strong>in</strong>ates<br />

Consider the <strong>in</strong>f<strong>in</strong>itessimal rectangle ∆∆ po<strong>in</strong>t<strong>in</strong>g <strong>in</strong> the k b direction shown <strong>in</strong> figure 4<br />

The l<strong>in</strong>e <strong>in</strong>tegral, taken <strong>in</strong> a right-handed way around k b gives<br />

I<br />

µ<br />

F · l = ∆ + + µ<br />

<br />

∆ − + <br />

µ<br />

<br />

∆ <br />

− ∆ =<br />

− <br />

<br />

∆∆<br />

<br />

<br />

Thus s<strong>in</strong>ce ∆∆ = ∆S the component of the curl is given by<br />

H µ<br />

(F) · bk <br />

= <br />

F · l<br />

∆S · bn = <br />

− <br />

<br />

(H.24)<br />

<br />

z<br />

F z<br />

(H.23)<br />

The same argument for the component of the curl <strong>in</strong> the direction<br />

is given by<br />

µ <br />

(F) ·bj =<br />

− <br />

<br />

(H.25)<br />

<br />

Similarly the same argument for the component of the curl <strong>in</strong> the <br />

direction is given by<br />

µ <br />

(F) ·bi =<br />

− <br />

<br />

(H.26)<br />

<br />

y<br />

x<br />

Figure H.4: Circulation around an<br />

<strong>in</strong>f<strong>in</strong>itessimal rectangle ∆∆ <strong>in</strong> the<br />

z direction


542 APPENDIX H. VECTOR INTEGRAL CALCULUS<br />

Thus comb<strong>in</strong><strong>in</strong>g the three components of the curl gives<br />

µ <br />

F =<br />

− <br />

<br />

µ<br />

b i +<br />

− <br />

<br />

Note that cross-product of the del operator with the vector F is<br />

b i b j k b<br />

∇ × F =<br />

<br />

<br />

¯ <br />

¯¯¯¯¯¯ <br />

µ<br />

b j +<br />

− <br />

bk<br />

<br />

(H.27)<br />

(H.28)<br />

which is identical to the right hand side of the relation for the curl <strong>in</strong> cartesian coord<strong>in</strong>ates. That is;<br />

∇ × F = −→ F<br />

(H.29)<br />

Therefore Stokes Theorem can be rewritten as<br />

I Z<br />

Z<br />

F · l = (F) · S =<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

(∇ × F) · S (H.30)<br />

The physics mean<strong>in</strong>g of the curl is that it is the circulation, or rotation, for an <strong>in</strong>f<strong>in</strong>itessimal loop at any<br />

location. The word curl is German for rotation.<br />

H.3 Example: Maxwell’s circulation equations<br />

As an example of the use of the curl, consider Faraday’s Law<br />

I<br />

Z<br />

E · l = −<br />

Us<strong>in</strong>g Stokes Theorem gives<br />

I<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Z<br />

E · l =<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

B<br />

· S<br />

(∇ × E) · S<br />

These two relations are <strong>in</strong>dependent of the shape of the closed loop, thus we obta<strong>in</strong> Faraday’s Law <strong>in</strong> the<br />

differential form<br />

(∇ × E) =− B<br />

<br />

Adifferential form of the Ampère-Maxwell law also can be obta<strong>in</strong>ed from<br />

I<br />

E<br />

B · l = 0<br />

Z(j + 0<br />

) · S<br />

Us<strong>in</strong>g Stokes Theorem<br />

I<br />

<br />

Z<br />

B · l =<br />

<br />

<br />

<br />

<br />

<br />

<br />

(∇ × B) · S<br />

Aga<strong>in</strong> this is <strong>in</strong>dependent of the shape of the loop and thus we obta<strong>in</strong><br />

Ampère-Maxwell law <strong>in</strong> differential form<br />

E<br />

∇ × B = 0 j + 0 0<br />

<br />

The differential forms of Maxwell’s circulation relations are easier to apply than the <strong>in</strong>tegral equations<br />

because the differential form relates the curl to the time derivatives at the same specific location.


H.4. POTENTIAL FORMULATIONS OF CURL-FREE AND DIVERGENCE-FREE FIELDS 543<br />

H.4 Potential formulations of curl-free and divergence-free fields<br />

Interest<strong>in</strong>g consequences result from the Divergence theorem and Stokes Theorem for vector fields that are<br />

either curl-free or divergence-free. In particular two theorems result from the second derivatives of a vector<br />

field.<br />

Theorem 1; Curl-free (irrotational) fields:<br />

For curl-free fields<br />

∇ × F =0<br />

(H.31)<br />

everywhere. This is automatically obeyed if the vector field is expressed as the gradient of a scalar field<br />

F = ∇<br />

(H.32)<br />

s<strong>in</strong>ce<br />

∇× (∇) =0 (H.33)<br />

That is, any curl-free vector field can be expressed <strong>in</strong> terms of the gradient of a scalar field.<br />

The scalar field is not unique, that is, any constant can be added to s<strong>in</strong>ce ∇ =0 that is, the<br />

addition of the constant does not change the gradient. This <strong>in</strong>dependence to addition of a number to the<br />

scalar potential is called a gauge <strong>in</strong>variance discussed <strong>in</strong> chapter 132 for which<br />

F = ∇ 0 = ∇ ( + ) =∇<br />

(H.34)<br />

That is, this gauge-<strong>in</strong>variant transformation does not change the observable F. The electrostatic field E<br />

and the gravitation field g are examples of irrotational fields that can be expressed as the gradient of scalar<br />

potentials.<br />

Theorem 2; Divergence-free (solenoidal) fields:<br />

For divergence-free fields<br />

∇ · F =0<br />

(H.35)<br />

everywhere. This is automatically obeyed if the field F is expressed <strong>in</strong> terms of the curl of a vector field G<br />

such that<br />

F = ∇ × G<br />

(H.36)<br />

s<strong>in</strong>ce ∇ · ∇ × G = 0. That is, any divergence-free vector field can be written as the curl of a related vector<br />

field.<br />

As discussed <strong>in</strong> chapter 132, the vector potential G is not unique <strong>in</strong> that a gauge transformation can be<br />

made by add<strong>in</strong>g the gradient of any scalar field, that is, the gauge transformation G 0 = G + ∇ϕ gives<br />

F = ∇ × G 0 = ∇× (G + ∇ϕ) =∇ × G<br />

(H.37)<br />

This gauge <strong>in</strong>variance for transformation to the vector potential G 0 does not change the observable vector<br />

field F The magnetic field B is an example of a solenoidal fieldthatcanbeexpressed<strong>in</strong>termsofthecurl<br />

of a vector potential A.<br />

H.4 Example: Electromagnetic fields:<br />

Electromagnetic <strong>in</strong>teractions are encountered frequently <strong>in</strong> classical mechanics so it is useful to discuss<br />

the use of potential formulations of electrodynamics.<br />

For electrostatics, Maxwell’s equations give that<br />

∇ × E =0<br />

Therefore theorem 1 states that it is possible to express this static electric field as the gradient of the scalar<br />

electric potential ,where<br />

E = −∇


544 APPENDIX H. VECTOR INTEGRAL CALCULUS<br />

For electrodynamics, Maxwell’s equations give that<br />

(∇ × E)+ B<br />

=0<br />

Assume that the magnetic field can be expressed <strong>in</strong> the terms of the vector potential B = ∇ × A, then<br />

the above equation becomes<br />

∇ × (E + A<br />

)=0<br />

Theorem 1 givesthatthiscurl-lessfield can be expressed as the gradient of a scalar field, here taken to<br />

be the electric potential .<br />

(E + A<br />

)==−∇<br />

that is<br />

E = −(∇ + A<br />

)<br />

Gauss’ law states that<br />

∇·E = 0<br />

which can be rewritten as<br />

∇·E = −∇ 2 (∇ · A)<br />

− = ()<br />

0<br />

Similarly <strong>in</strong>sertion of the vector potential A <strong>in</strong> Ampère’s Law gives<br />

µ µ<br />

E<br />

<br />

2 <br />

∇ × B = ∇ × (∇ × A)= 0 j + 0 0<br />

= A<br />

0j− 0 0 ∇ − <br />

0 0<br />

2<br />

Us<strong>in</strong>g the vector identity ∇ × (∇ × A) =∇ (∇ · A) −∇ 2 allows the above equation to be rewritten as<br />

µ<br />

µ <br />

∇ 2 2 µ<br />

µ <br />

A<br />

<br />

A− 0 0<br />

2 − ∇ ∇ · A+ 0 0 = −<br />

<br />

0 j ( )<br />

Theuseofthescalarpotential and vector potential A leads to two coupled equations and . These<br />

coupled equations can be transformed <strong>in</strong>to two uncoupled equations by exploit<strong>in</strong>g the freedom to make a gauge<br />

transformation for the vector potential such that the middle brackets <strong>in</strong> both equations and are zero.<br />

That is, choos<strong>in</strong>g the Lorentz gauge<br />

µ <br />

∇ · A = − 0 0<br />

<br />

simplifies equations and to be<br />

∇ 2 2 <br />

− 0 0<br />

2 = − <br />

µ 0<br />

<br />

∇ 2 2 <br />

A<br />

A− 0 0<br />

2 = − 0 j<br />

The virtue of us<strong>in</strong>g the Lorentz gauge, rather than the Coulomb gauge ∇ · A =0 is that it separates the<br />

equations for the scalar and vector potentials. Moreover, these two equations are the wave equations for these<br />

two potential fields correspond<strong>in</strong>g to a velocity = √ 1<br />

0 0<br />

. This example illustrates the power of us<strong>in</strong>g the<br />

concept of potentials <strong>in</strong> describ<strong>in</strong>g vector fields.


Appendix I<br />

Waveform analysis<br />

I.1 Harmonic waveform decomposition<br />

Any l<strong>in</strong>ear system that is subject to a time-dependent forc<strong>in</strong>g function () can be expressed as a l<strong>in</strong>ear<br />

superposition of frequency-dependent solutions of the <strong>in</strong>dividual harmonic decomposition () of the forc<strong>in</strong>g<br />

function. Similarly, any l<strong>in</strong>ear system subject to a spatially-dependent forc<strong>in</strong>g function () can be expressed<br />

as a l<strong>in</strong>ear superposition of the wavenumber-dependent solutions of the <strong>in</strong>dividual harmonic decomposition<br />

( ) of the forc<strong>in</strong>g function. Fourier analysis provides the mathematical procedure for the transformation<br />

between the periodic waveforms and the harmonic content, that is, () ⇔ (), or () ⇔ ( ).Fourier’s<br />

theorem states that any arbitrary forc<strong>in</strong>g function () can be decomposed <strong>in</strong>to a sum of harmonic terms.<br />

For example for a time-dependent periodic forc<strong>in</strong>g function the decomposition can be a cos<strong>in</strong>e series of the<br />

form<br />

∞X<br />

() = cos( 0 + )<br />

(I.1)<br />

=1<br />

where 0 is the lowest (fundamental) frequency solution. For an aperiodic function a cos<strong>in</strong>e decomposition<br />

can be of the form<br />

Z ∞<br />

() = ()cos( + ())<br />

(I.2)<br />

0<br />

Either of the complementary functions () ⇔ (), or () ⇔ ( ) are equivalent representations of<br />

the harmonic content that can be used to describe signals and waves. The follow<strong>in</strong>g two sections give an<br />

<strong>in</strong>troduction to Fourier analysis.<br />

I.1.1<br />

Periodic systems and the Fourier series<br />

Discrete solutions occur for systems when periodic boundary conditions exist. The response of periodic<br />

systems can be described <strong>in</strong> either the time versus angular frequency doma<strong>in</strong>s, or equivalently, the spatial<br />

coord<strong>in</strong>ate versus the correspond<strong>in</strong>g wave number . For periodic systems this decomposition leads to<br />

the Fourier series where a generalized phase coord<strong>in</strong>ate can be used to represent either the time or spatial<br />

coord<strong>in</strong>ates, that is, with = 0 or = respectively. The Fourier series relates the two representations<br />

ofthediscretewavesolutionsforsuchperiodicsystems.<br />

Fourier’s theorem states that for a general periodic system any arbitrary forc<strong>in</strong>g function () can be<br />

decomposed <strong>in</strong>to a sum of s<strong>in</strong>usoidal or cos<strong>in</strong>usoidal terms. The summation can be represented by three<br />

equivalent series expansions given below, where = 0 or = k 0·r and where 0 k 0 are the fundamental<br />

angular frequency and fundamental wave number respectively.<br />

() = ∞ 0<br />

2 + X<br />

[ cos ()+ s<strong>in</strong> ()] (I.3)<br />

=1<br />

() = ∞ 0<br />

2 + X<br />

cos ( + )<br />

=0<br />

545<br />

(I.4)


546 APPENDIX I. WAVEFORM ANALYSIS<br />

() = ∞<br />

0<br />

2 + X<br />

s<strong>in</strong> ( + )<br />

(I.5)<br />

=0<br />

where is an <strong>in</strong>teger, and are phase shifts fit to the <strong>in</strong>itial conditions.<br />

The normal modes of a discrete system form a complete set of solutions that satisfy the follow<strong>in</strong>g orthogonality<br />

relation<br />

Z 2<br />

() () = <br />

(I.6)<br />

0<br />

where is the Kronecker delta symbol def<strong>in</strong>ed<strong>in</strong>equation(10). Orthogonality can be used to determ<strong>in</strong>e<br />

the coefficients for equations (3) to be<br />

0 = 1 <br />

= 1 <br />

= 1 <br />

Z +<br />

−<br />

Z +<br />

−<br />

Z +<br />

−<br />

() <br />

()cos() <br />

()s<strong>in</strong>() <br />

(I.7)<br />

(I.8)<br />

(I.9)<br />

Similarly the coefficients for (4) and (5) are related to the above coefficients by<br />

2 = 2 = 2 + 2 <br />

Instead of the simple trigonometric form used <strong>in</strong> equations (3 − 5) the cos<strong>in</strong>e and s<strong>in</strong>e functions can<br />

be expanded <strong>in</strong>to the exponential form where<br />

then equation (3) becomes<br />

cos = 1 ¡<br />

+ −¢ (I.10)<br />

2<br />

s<strong>in</strong> = − ¡<br />

− −¢<br />

2<br />

() =<br />

∞X<br />

=−∞<br />

<br />

where is any <strong>in</strong>teger and, from the orthogonality, the Fourier coefficients are given by<br />

= 1<br />

2<br />

Z +<br />

−<br />

() <br />

(I.11)<br />

(I.12)<br />

These coefficients are related to the cos<strong>in</strong>e plus s<strong>in</strong>e series amplitudes by<br />

= 1 2 ( − ) (when is positive)<br />

= 1 2 ( + ) (when is negative)<br />

These results show that the coefficients of the exponential series are <strong>in</strong> general complex, and that they<br />

occur <strong>in</strong> conjugate pairs (that is, the imag<strong>in</strong>ary part of a coefficient is equal but opposite <strong>in</strong> sign to that<br />

for the coefficient − ). Although the <strong>in</strong>troduction of complex coefficients may appear unusual, it should<br />

be remembered that the real part of a pair of coefficients denotes the magnitude of the cos<strong>in</strong>e wave of the<br />

relevant frequency, and that the imag<strong>in</strong>ary part denotes the magnitude of the s<strong>in</strong>e wave. If a particular<br />

pair of coefficients and − are real, then the component at the frequency 0 is simply a cos<strong>in</strong>e; if <br />

and − are purely imag<strong>in</strong>ary, the component is just a s<strong>in</strong>e; and if, as is the general case, and − are<br />

complex, both cos<strong>in</strong>e and a s<strong>in</strong>e terms are present.<br />

The use of the exponential form of the Fourier series gives rise to the notion of ‘negative frequency’. Of<br />

course, () = cos is a wave of a s<strong>in</strong>gle frequency = 0 radians/second, and may be represented


I.1. HARMONIC WAVEFORM DECOMPOSITION 547<br />

by a s<strong>in</strong>gle l<strong>in</strong>e of height <strong>in</strong> a normal spectral diagram. However, us<strong>in</strong>g the exponential form of the Fourier<br />

series results <strong>in</strong> both positive and negative components.<br />

The coexistence of both negative and positive angular frequencies ± can be understood by consideration<br />

of the Argand diagram where the real component is plotted along the -axis and the imag<strong>in</strong>ary component<br />

along the -axis. The function + represents a vector of length that rotates with an angular velocity <br />

<strong>in</strong> a positive direction, that is counterclockwise, whereas, − represents the vector rotat<strong>in</strong>g <strong>in</strong> a negative<br />

direction, that is clockwise. Thus the sum of the two rotat<strong>in</strong>g vectors, accord<strong>in</strong>g to equations (3), leads<br />

to cancellation of the opposite components on the imag<strong>in</strong>ary axis and addition of the two cos real<br />

components on the axis. Subtraction leads to cancellation of the real components and addition of the<br />

imag<strong>in</strong>ary axis components.<br />

I.1.2<br />

Aperiodic systems and the Fourier Transform<br />

The Fourier transform (also called the Fourier <strong>in</strong>tegral) does for the non-repetitive signal waveform what<br />

the Fourier series does for the repetitive signal. It was shown that the l<strong>in</strong>e spectrum of a recurrent periodic<br />

pulse waveform is modified as the pulse duration decreases, assum<strong>in</strong>g the period of the waveform (and hence<br />

its fundamental component) rema<strong>in</strong>s unchanged. Suppose now that the duration of the pulses rema<strong>in</strong> fixed<br />

but the separation between them <strong>in</strong>creases, giv<strong>in</strong>g rise to an <strong>in</strong>creas<strong>in</strong>g period. In the limit, only a s<strong>in</strong>gle<br />

rectangular pulse rema<strong>in</strong>s, its neighbors hav<strong>in</strong>g moved away on either side towards ±∞. In this case, the<br />

fundamental frequency 0 tends towards zero and the harmonics become extremely closely spaced and of<br />

vanish<strong>in</strong>gly small amplitudes, that is, the system approximates a cont<strong>in</strong>uous spectrum.<br />

Mathematically, this situation may be expressed by modifications to the exponential form of the Fourier<br />

series already derived. Let the phase factor = 0 <strong>in</strong> equation (11) then<br />

= 0<br />

2<br />

Z +<br />

−<br />

() 0 = 1 <br />

Z + <br />

2<br />

− 2<br />

() 0 <br />

(I.13)<br />

where is the period of the periodic force. Let () = , = 0 and take the limit for →∞ then<br />

equation (12) can be written as<br />

Z +∞<br />

() = () <br />

(I.14)<br />

−∞<br />

Similarly mak<strong>in</strong>g the same limit for →∞then 0 = 2 <br />

→ and equation (11) becomes<br />

() =<br />

∞X<br />

=−∞<br />

()<br />

0 =<br />

<br />

∞X<br />

=−∞<br />

() 0<br />

2 = 1<br />

2<br />

Z +∞<br />

−∞<br />

() <br />

(I.15)<br />

Equation (15) shows how a non-repetitive time-doma<strong>in</strong> wave form is related to its cont<strong>in</strong>uous spectrum.<br />

These are known as Fourier <strong>in</strong>tegrals or Fourier transforms. They are of central importance for signal<br />

process<strong>in</strong>g. For convenience the transforms often are written <strong>in</strong> the operator formalism us<strong>in</strong>g the F symbol<br />

<strong>in</strong> the form<br />

() = 1 Z +∞<br />

∙ ()¸ 1<br />

() ≡ F −1 (I.16)<br />

2<br />

2<br />

() =<br />

−∞<br />

Z +∞<br />

−∞<br />

() − ≡ F()<br />

(I.17)<br />

It is very important to grasp the significance of these two equations. The first tells us that the Fourier<br />

transform of the waveform () is cont<strong>in</strong>uously distributed <strong>in</strong> the frequency range between = ±∞, whereas<br />

the second shows how, <strong>in</strong> effect, the waveform may be synthesized from an <strong>in</strong>f<strong>in</strong>ite set of exponential functions<br />

of the form ± , each weighted by the relevant value of (). It is crucial to realize that this transformation<br />

can go either way equally, that is, from () to () or vice versa. 1<br />

1 The only asymmetry <strong>in</strong> the Fourier transform relations comes from the 2 factor orig<strong>in</strong>at<strong>in</strong>g from the fact that by convention<br />

physicists use the angular frequency =2 rather than the frequency . In order to restore symmetry many papers use the<br />

1<br />

factor √ <strong>in</strong> both relations rather than us<strong>in</strong>g the 1 factor <strong>in</strong> equation 16 and unity <strong>in</strong> equation 17.<br />

2<br />

2


548 APPENDIX I. WAVEFORM ANALYSIS<br />

I.1 Example: Fourier transform of a s<strong>in</strong>gle isolated square pulse:<br />

as<br />

Consider a s<strong>in</strong>gle isolated square pulse of width that is described by the rectangular function Π def<strong>in</strong>ed<br />

½ 1 || <br />

Π() =<br />

2<br />

0 || 2<br />

That is, assume that the amplitude of the pulse is unity between − 2 ≤ ≤ 2<br />

() =<br />

Z +<br />

−<br />

1 − = <br />

µ s<strong>in</strong><br />

<br />

2<br />

<br />

2<br />

<br />

. Then the Fourier transform<br />

which is an unnormalized () function. Note that the width of the pulse ∆ = ± 2<br />

leads to a frequency<br />

envelope that has the first zeros at ∆ = ± <br />

. Thus the product of these widths ∆ · ∆ = ± which is<br />

<strong>in</strong>dependent of the width of the pulse, that is ∆ = ∆<br />

which is an example of the uncerta<strong>in</strong>ty pr<strong>in</strong>ciple<br />

which is applicable to all forms of wave motion.<br />

I.2 Example: Fourier transform of the Dirac delta function:<br />

The Dirac delta function, ( − 0 ), is a pulse of extremely short duration and unit area at = 0 and is<br />

zero at all other times. That is,<br />

1=<br />

Z +∞<br />

−∞<br />

( − 0 ) <br />

The Dirac function, which is sometimes referred to as the impulse function, has many important applications<br />

to physics and signal process<strong>in</strong>g. For example, a shell shot from a gun is given a mechanical impulse<br />

impart<strong>in</strong>g a certa<strong>in</strong> momentum to the shell <strong>in</strong> a very short time. Other th<strong>in</strong>gs be<strong>in</strong>g equal, one is <strong>in</strong>terested<br />

only <strong>in</strong> the impulse imparted to the shell, that is, the time <strong>in</strong>tegral of the force accelerat<strong>in</strong>g the shell <strong>in</strong> the<br />

gun, rather than the details of the time dependence of the force. S<strong>in</strong>ce the force acts for a very short time<br />

the Dirac delta function can be employed <strong>in</strong> such problems.<br />

As described <strong>in</strong> section 311 and appendix , the Dirac delta function is employed <strong>in</strong> signal process<strong>in</strong>g<br />

when signals are sampled for short time <strong>in</strong>tervals. The Fourier transform of the delta function is needed for<br />

discussion of sampl<strong>in</strong>g of signals<br />

() =<br />

Z +∞<br />

−∞<br />

( − 0 ) − = −0<br />

S<strong>in</strong>ce − essentially is constant over the <strong>in</strong>f<strong>in</strong>itesimal time duration of the ( − 0 ) function, and the<br />

time <strong>in</strong>tegral of the function is unity, thus the term − hasunitmagnitudeforanyvalueof and has<br />

a phase shift of − ( − 0 )radians. For 0 =0the phase shift is zero and thus the Fourier transform of a<br />

Dirac () function is () =1. That is, this is a uniform white spectrum for all values of .<br />

I.2 Time-sampled waveform analysis<br />

An alternative approach for unloos<strong>in</strong>g periodic signals, that is complementary to the Fourier analysis harmonic<br />

decomposition, is time-sampled (discrete-sample) waveform analysis where the signal amplitude is<br />

measured repetitively at regular time <strong>in</strong>tervals <strong>in</strong> a time-ordered sequence, that is, a sequence of samples of<br />

the <strong>in</strong>stantaneous delta-function amplitudes is recorded. Typically an amplitude-to-digital converter is used<br />

to digitize the amplitude for each measured sample and the digital numbers are recorded; this process is<br />

called digital signal process<strong>in</strong>g.<br />

The general pr<strong>in</strong>ciples are best expla<strong>in</strong>ed by first consider<strong>in</strong>g the response of a l<strong>in</strong>ear system to a step<br />

function impulse, followed by a square impulse, and lead<strong>in</strong>g to the response of a -function impulsive driv<strong>in</strong>g<br />

force.


I.2. TIME-SAMPLED WAVEFORM ANALYSIS 549<br />

Figure I.1: Response of a underdamped l<strong>in</strong>ear oscillator with =10,andΓ =2to the follow<strong>in</strong>g impulsive<br />

force. (a) Step function force =0for 0 and = for 0 (b) Square-wave force where = for<br />

0 for =3 and =0at other times. (c) Delta-function impulse =1.<br />

I.2.1<br />

Delta-function impulse response<br />

Consider the damped oscillator equation<br />

¨ + Γ ˙ + 2 0 = ()<br />

<br />

and assume that a step function is applied at time =0.Thatis;<br />

(I.18)<br />

()<br />

=0 0 ()<br />

= 0 (I.19)<br />

where is a constant. The <strong>in</strong>itial conditions are that (0) = ˙(0) = 0.<br />

The transient or complementary solution is the solution of the l<strong>in</strong>early-damped harmonic oscillator<br />

¨ + Γ ˙ + 2 0 =0<br />

(I.20)<br />

This is <strong>in</strong>dependent of the driv<strong>in</strong>g force and the solution is given <strong>in</strong> the chapter 35 discussion of the l<strong>in</strong>earlydamped<br />

harmonic oscillator.<br />

The particular, steady-state, solution is easy to obta<strong>in</strong> just by <strong>in</strong>spection s<strong>in</strong>ce the force is a constant,<br />

that is, the particular solution is<br />

= 2 0<br />

0 =0 0<br />

Tak<strong>in</strong>g the sum of the transient and particular solutions, us<strong>in</strong>g the <strong>in</strong>itial conditions, gives the f<strong>in</strong>al solution<br />

to be<br />

"<br />

#<br />

() = 2 1 − − Γ 2 cos 1 − Γ− Γ 2 <br />

s<strong>in</strong> 1 <br />

(I.21)<br />

0<br />

2 1<br />

q<br />

where 1 ≡ 2 0 − ¡ ¢<br />

Γ 2<br />

2 This functional form is shown <strong>in</strong> figure 1. Note that the amplitude of the<br />

transient response equals − at =0to cancel the particular solution when it jumps to +. The oscillatory<br />

behavior then is just that of the transient response.<br />

A square impulse can be generated by the superposition of two opposite-sign stepfunctions separated by<br />

atime as shown <strong>in</strong> figure 1.<br />

The square impulse can be taken to the limit where the width is negligibly small relative to the response<br />

times of the system. It can be shown that lett<strong>in</strong>g → 0 but keep<strong>in</strong>g the magnitude of the total impulse<br />

= f<strong>in</strong>ite for the impulse at time 0 , leads to the solution for the -function impulse occurr<strong>in</strong>g at 0<br />

() = 1<br />

− Γ 2 (−0) s<strong>in</strong> 1 ( − 0 ) 0 (I.22)<br />

This response to a delta function impulse is shown <strong>in</strong> figure 1 for the case where 0 =0. An example is<br />

the response when the hammer strikes a piano str<strong>in</strong>g at =0.


550 APPENDIX I. WAVEFORM ANALYSIS<br />

Figure I.2: Decomposition of the function () =2s<strong>in</strong>()+s<strong>in</strong>(5)+ 1 3 s<strong>in</strong> (15)+ 1 5<br />

s<strong>in</strong>(25) <strong>in</strong>to a time-ordered<br />

sequence of -function samples.<br />

I.2.2<br />

Green’s function waveform decomposition<br />

The response of the l<strong>in</strong>early-damped l<strong>in</strong>ear oscillator to an delta function impulse, that has been expressed<br />

above, can be used to exploit the powerful Green’s technique for decomposition of any general forc<strong>in</strong>g<br />

function. That is, if the driven system is l<strong>in</strong>ear, then the pr<strong>in</strong>ciple of superposition is applicable and allow<strong>in</strong>g<br />

expression of the <strong>in</strong>homogeneous part of the differential equation as the sum of <strong>in</strong>dividual delta functions.<br />

That is;<br />

¨ + Γ ˙ + 2 0 =<br />

∞X<br />

=−∞<br />

()<br />

= ∞ X<br />

=−∞<br />

()<br />

(I.23)<br />

As illustrated <strong>in</strong> figure 2 discrete-time waveform analysis <strong>in</strong>volves repeatedly sampl<strong>in</strong>g the <strong>in</strong>stantaneous<br />

amplitude <strong>in</strong> a regular and repetitive sequence of -function impulses. S<strong>in</strong>ce the superposition pr<strong>in</strong>ciple<br />

applies for this l<strong>in</strong>ear system then the waveform can be described by a sum of an ordered series of deltafunction<br />

impulses where 0 is the time of an impulse. Integrat<strong>in</strong>g over all the -function responses that have<br />

occurred at time 0 , that is prior to the time of <strong>in</strong>terest leads to<br />

() =<br />

Z <br />

−∞<br />

The Green’s function ( − 0 ) is def<strong>in</strong>ed by<br />

( 0 )<br />

1<br />

− Γ 2 (− 0 ) s<strong>in</strong> 1 ( − 0 ) 0 ≥ 0 (I.24)<br />

( − 0 1<br />

) = − Γ 2 (− 0 ) s<strong>in</strong> 1 ( − 0 ) ≥ 0 (I.25)<br />

1<br />

= 0 0<br />

Superposition allows the summed response of the system to be written <strong>in</strong> an <strong>in</strong>tegral form<br />

() =<br />

Z <br />

−∞<br />

( 0 )( − 0 ) 0<br />

(I.26)<br />

which gives the f<strong>in</strong>al time dependence of the forced system. This repetitive time-sampl<strong>in</strong>g approach avoids<br />

the need of us<strong>in</strong>g Fourier analysis. Note that the Green’s function ( − 0 ) <strong>in</strong>cludes<br />

q<br />

implicitly the frequency<br />

of the free undamped l<strong>in</strong>ear oscillator 0 the free damped l<strong>in</strong>ear oscillator 1 ≡ 2 0 − ¡ ¢<br />

Γ 2<br />

2 as well as the<br />

damp<strong>in</strong>g coefficient Γ. Access to the comb<strong>in</strong>ation of fast microcomputers coupled to fast digital sampl<strong>in</strong>g<br />

techniques has made digital signal sampl<strong>in</strong>g the pre-em<strong>in</strong>ent technique for signal record<strong>in</strong>g of audio, video,<br />

and detector signal process<strong>in</strong>g.


Bibliography<br />

[1] SELECTION OF TEXTBOOKS ON CLASSICAL MECHANICS<br />

[Ar78] V. I. Arnold, “Mathematical methods of <strong>Classical</strong> <strong>Mechanics</strong>”, 2 edition, Spr<strong>in</strong>ger-Verlag (1978)<br />

This textbook provides an elegant and advanced exposition of classical mechanics expressed <strong>in</strong><br />

the language of differential topology.<br />

[Co50] H.C. Corben and P. Stehle, “<strong>Classical</strong> <strong>Mechanics</strong>”, John Wiley (1950)<br />

This classic textbook covers the material at the same level and comparable scope as the present<br />

textbook.<br />

[Fo05] G. R. Fowles, G. L. Cassiday, “Analytical <strong>Mechanics</strong>”. Thomson Brookes/Cole, Belmont, (2005)<br />

An elementary undergraduate text that emphasizes computer simulations.<br />

[Go50] H. Goldste<strong>in</strong>, “<strong>Classical</strong> <strong>Mechanics</strong>”, Addison-Wesley, Read<strong>in</strong>g (1950)<br />

This has rema<strong>in</strong>ed the gold standard graduate textbook <strong>in</strong> classical mechanics s<strong>in</strong>ce 1950. Goldste<strong>in</strong>’s<br />

book is the best graduate-level reference to supplement the present textbook. The lack of<br />

worked examples is an impediment to us<strong>in</strong>g Goldste<strong>in</strong> for undergraduate courses. The 3 edition,<br />

published by Goldste<strong>in</strong>, Poole, and Safko (2002), uses the symplectic notation that makes the<br />

book less friendly to undergraduates. The Cl<strong>in</strong>e book adopts the nomenclature used by Goldste<strong>in</strong><br />

to provide a consistent presentation of the material.<br />

[Gr06]<br />

[Gr10]<br />

R. D. Gregory, “<strong>Classical</strong> <strong>Mechanics</strong>”, Cambridge University Press<br />

This outstand<strong>in</strong>g, and orig<strong>in</strong>al, <strong>in</strong>troduction to analytical mechanics was written by a mathematician.<br />

It is ideal for the undergraduate, but the breadth of the material covered is limited.<br />

W. Gre<strong>in</strong>er, “<strong>Classical</strong> <strong>Mechanics</strong>, Systems of particles and Hamiltonian Dynamics” ,2 edition,<br />

Spr<strong>in</strong>ger (2010). This excellent modern graduate textbook is similar <strong>in</strong> scope and approach to<br />

the present text. Gre<strong>in</strong>er <strong>in</strong>cludes many <strong>in</strong>terest<strong>in</strong>g worked examples, as well as a reproduction<br />

of the Struckmeier[Str08] presentation of the extended Lagrangian and Hamiltonian mechanics<br />

formalism of Lanczos[La49].<br />

[Jo98] J. V. José and E. J. Saletan, “<strong>Classical</strong> Dynamics, A Contemporary Approach”, Cambridge<br />

University Press (1998)<br />

This modern advanced graduate-level textbook emphasizes configuration manifolds and tangent<br />

bundles which makes it unsuitable for use by most undergraduate students.<br />

[Jo05]<br />

O. D. Johns, “Analytical <strong>Mechanics</strong> for Relativity and Quantum <strong>Mechanics</strong>”, 2 edition, Oxford<br />

University Press (2005). Excellent modern graduate text that emphasizes the Lanczos[La49]<br />

parametric approach to Special Relativity. The Johns and Cl<strong>in</strong>e textbooks were developed <strong>in</strong>dependently<br />

but are similar <strong>in</strong> scope and approach. For consistency, the name “generalized energy”,<br />

which was <strong>in</strong>troduced by Johns, has been adopted <strong>in</strong> the Cl<strong>in</strong>e textbook.<br />

551


552 BIBLIOGRAPHY<br />

[Ki85]<br />

[La49]<br />

[La60]<br />

T.W.B. Kibble, F.H. Berkshire. “<strong>Classical</strong> <strong>Mechanics</strong>, (5th edition)”, Imperial College Press,<br />

London, 2004. Based on the textbook written by Kibble that was published <strong>in</strong> 1966 by McGraw-<br />

Hill. The 4th and 5th editions were published jo<strong>in</strong>tly by Kibble and Berkshire. This excellent<br />

and well-established textbook addresses the same undergraduate student audience as the present<br />

textbook. This book covers the variational pr<strong>in</strong>ciples and applications with m<strong>in</strong>imal discussion of<br />

the philosophical implications of the variational approach.<br />

C. Lanczos, “The <strong>Variational</strong> <strong>Pr<strong>in</strong>ciples</strong> of <strong>Mechanics</strong>”, University of Toronto Press, Toronto,<br />

(1949)<br />

An outstand<strong>in</strong>g graduate textbook that has been one of the found<strong>in</strong>g pillars of the field s<strong>in</strong>ce<br />

1949. It gives an excellent <strong>in</strong>troduction to the philosophical aspects of the variational approach<br />

to classical mechanics, and <strong>in</strong>troduces the extended formulations of Lagrangian and Hamiltonian<br />

mechanics that are applicable to relativistic mechanics.<br />

L. D. Landau, E. M. Lifshitz, “<strong>Mechanics</strong>”, Volume 1 of a Course <strong>in</strong> Theoretical Physics, Pergamon<br />

Press (1960)<br />

An outstand<strong>in</strong>g, succ<strong>in</strong>ct, description of analytical mechanics that is devoid of any superfluous<br />

text. This Course <strong>in</strong> Theoretical Physics is a masterpiece of scientific writ<strong>in</strong>gandisanessential<br />

component of any physics library. The compactness and lack of examples makes this textbook<br />

less suitable for most undergraduate students.<br />

[Li94] Yung-Kuo Lim, “Problems and Solutions on <strong>Mechanics</strong>” (1994)<br />

This compendium of 408 solved problems, which are taken from graduate qualify<strong>in</strong>g exam<strong>in</strong>ations<br />

<strong>in</strong> physics at several U.S. universities, provides an <strong>in</strong>valuable resource that complements this<br />

textbook for study of Lagrangian and Hamiltonian mechanics.<br />

[Ma65] J. B. Marion, “<strong>Classical</strong> Dynamics of Particles and Systems”, Academic Press, New York, (1965)<br />

This excellent undergraduate text played a major role <strong>in</strong> <strong>in</strong>troduc<strong>in</strong>g analytical mechanics to<br />

the undergraduate curriculum. It has an outstand<strong>in</strong>g collection of challeng<strong>in</strong>g problems. The 5 <br />

edition has been published by S. T. Thornton and J. B. Marion, Thomson, Belmont, (2004).<br />

[Me70] L. Meirovitch, “Methods of Analytical Dynamics”, McGraw-Hill New York, (1970)<br />

An advanced eng<strong>in</strong>eer<strong>in</strong>g textbook that emphasizes solv<strong>in</strong>g practical problems, rather than the<br />

underly<strong>in</strong>g theory.<br />

[Mu08] H. J. W. Müller-Kirsten, “<strong>Classical</strong> <strong>Mechanics</strong> and Relativity”, World Scientific, S<strong>in</strong>gapore, (2008)<br />

This modern graduate-level textbook emphasizes relativistic mechanics mak<strong>in</strong>g it an excellent<br />

complement to the present textbook.<br />

[Pe82]<br />

[Sy60]<br />

I. Percival and D. Richards, “Introduction to Dynamics” Cambridge University Press, London,<br />

(1982)<br />

Provides a clear presentation of Lagrangian and Hamiltonian mechanics, <strong>in</strong>clud<strong>in</strong>g canonical<br />

transformations, Hamilton-Jacobi theory, and action-angle variables.<br />

J.L. Synge, “<strong>Pr<strong>in</strong>ciples</strong> of <strong>Classical</strong> <strong>Mechanics</strong> and Field Theory” , Volume III/I of “Handbuck<br />

der Physik” Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong> (1960).<br />

A classic graduate-level presentation of analytical mechanics.<br />

[Ta05] J. R. Taylor, “<strong>Classical</strong> <strong>Mechanics</strong>”, University Science Books, Sausalito, (2006)<br />

This undergraduate book gives a well-written descriptive <strong>in</strong>troduction to analytical mechanics.<br />

The scope of the book is limited and the problems are easy.


BIBLIOGRAPHY 553<br />

[2] GENERAL REFERENCES<br />

[Bak96]<br />

L. Baker, J.P. Gollub, Chaotic Dynamics, 2 edition, 1996 (Cambridge University Press)<br />

[Bat31] H. Bateman, Phys. Rev. 38 (1931) 815<br />

[Bau31] P.S. Bauer, Proc. Natl. Acad. Sci. 17 (1931) 311<br />

[Bor25a] M. Born and P. Jordan, Zur Quantenmechanik, Zeitschrift für Physik, 34, (1925) 858-888.<br />

[Bor25b] M. Born, W. Heisenberg, and P. Jordan, Zur Quantenmechanik II, Zeitschrift für Physik, 35,<br />

(1925), 557-615,<br />

[Boy08]<br />

R. W. Boyd, Nonl<strong>in</strong>ear Optics, 3 edition, 2008 (Academic Press, NY)<br />

[Bri14] L. Brillou<strong>in</strong>, Ann. Physik 44(1914)<br />

[Bri60]<br />

L. Brillou<strong>in</strong>, Wave Propagation and Group Velocity, 1960 (Academic Press, New York)<br />

[Cay1857] A. Cayley, Proc. Roy. Soc. London 8 (1857) 506<br />

[Cei10] J.L. Cieśliński, T. Nikiciuk, J. Phys. A:Math. Theor. 43 (2010) 175205<br />

[Cio07] Ciocci and Langerock, Regular and Chaotic Dynamics, 12 (2007) 602<br />

[Cli71] D. Cl<strong>in</strong>e, Proc. Orsay Coll. on Intermediate Nuclei, Ed. Foucher, Perr<strong>in</strong>, Veneroni, 4 (1971).<br />

[Cli72]<br />

D. Cl<strong>in</strong>e and C. Flaum, Proc. of the Int. Conf. on Nuclear Structure Studies Us<strong>in</strong>g Electron<br />

Scatter<strong>in</strong>g, Sendai, Ed. Shoa, Ui, 61 (1972).<br />

[Cli86] D. Cl<strong>in</strong>e, Ann. Rev. Nucl. Part. Sci. 36, (1986) 683.<br />

[Coh77] R.J. Cohen, Amer. J. of Phys. 45 (1977) 12<br />

[Cra65]<br />

F.S. Crawford, Berkeley Physics Course 3; Waves, 1970 (Mc Graw Hill, New York)<br />

[Cum07] D. Cum<strong>in</strong>, C.P. Unsworth, Physica D 226 (2007) 181<br />

[Dav58] A. S. Davydov and G. F. Filippov. Nuclear Physics, 8 (1958) 237<br />

[Dek75] H. Dekker, Z. Physik, B21 (1975) 295<br />

[Dep67] A. Deprit, American J. of Phys 35, no.5 424 (1967)<br />

[Dir30] P.A.M. Dirac, Quantum <strong>Mechanics</strong>, Oxford University Press, (1930).<br />

[Dou41] D. Douglas, Trans. Am. Math. Soc. 50 (1941) 71<br />

[Fey84]<br />

R.P. Feynman, R.B. Leighton, M. Sands, The Feynman Lectures, (Addison-Wesley, Read<strong>in</strong>g,<br />

MA,1984) Vol. 2, p17.5<br />

[Fro80] C. Frohlich, Scientific American,242 (1980) 154<br />

[Gal13] C. R. Galley, Physical Review Letters, 11 (2013) 174301<br />

[Gal14] C. R. Galley, D. Tsang, L.C. Ste<strong>in</strong>, arXiv:1412.3082v1 [math-phys] 9 Dec 2014<br />

[Har03]<br />

James B. Hartle, Gravity: An Introduction to E<strong>in</strong>ste<strong>in</strong>’s General Relativity (Addison Wesley,<br />

2003)<br />

[Jac75] J.D. Jackson, <strong>Classical</strong> Electrodynamics, 2 edition , (Wiley, 1975)<br />

[Kur75]<br />

International Symposium on Math. Problems <strong>in</strong> Theoretical Physics, Lecture Notes <strong>in</strong> Physics,<br />

Vol39 Spr<strong>in</strong>ger, NY (1975)<br />

[Mus08a] Z.E. Musielak, J. Phys. A. Math. Theor. 41 (2008) 055205


554 BIBLIOGRAPHY<br />

[Mus08b] Z.E. Musielak, D. Rouy, L.D. Swift, Chaos, Solitons, Fractals 38 (2008) 894<br />

[Ray1881] J.W. Strutt, 3 Baron Rayleigh, Proc. London Math. Soc., s1-4 (1), (1881) 357<br />

[Ray1887] J.W. Strutt, 3 Baron Rayleigh, The Theory of Sound, 1887 (Macmillan, London)<br />

[Rou1860] E.J. Routh, Treatise on the dynamics of a system of rigid bodies, MacMillan (1860)<br />

[Sim98]<br />

M. Simon, D. Cl<strong>in</strong>e, K. Vetter, et al, Unpublished<br />

[Sta05] T. Stachowiak and T. Okada, Chaos, Solitons, and Fractals, 29 (2006) 417.<br />

[Str00] S.H. Strogatz, Physica D43 (2000) 1<br />

[Str05] J. Struckmeier, J. Phys. A: Math; Gen. 38 (2005) 1257<br />

[Str08] J. Struckmeier, Int. J. of Mod. Phys. E18 (2008) 79<br />

[Vir15] E.G. Virga, Phys, Rev. E91 (2015) 013203<br />

[W<strong>in</strong>67] A.T. W<strong>in</strong>free, J. Theoretical Biology 16 (1967) 15


Index<br />

Abbreviated action, 224<br />

Action<br />

abbreviated action, 224<br />

Hamilton’s Action Pr<strong>in</strong>ciple, 222<br />

Action-angle variables<br />

Hamilton-Jacobi theory, 425<br />

Sommerfeld atom, 487<br />

Adiabatic <strong>in</strong>variance<br />

action variables, 428<br />

plane pendulum, 428<br />

Analytical mechanics, xviii<br />

Androyer-Deprit variables<br />

rigid-body rotation, 333<br />

Archimedes<br />

history, 1<br />

Aristotle<br />

history, 1<br />

Asymmetric rotor<br />

stability of torque-free rotation, 340<br />

Asymmetric top<br />

5 somersaults plus 3 rotations of high diver, 351<br />

separatrix, 340<br />

tennis racket motion, 341<br />

torque-free rotation, 339<br />

Attractor<br />

van der Pol oscillator, 90<br />

Autonomous system, 89, 165<br />

Barycenter, 247<br />

Bernoulli<br />

history, 5<br />

pr<strong>in</strong>ciple of virtual work, 134<br />

virtual work, 107<br />

Bertrand’s Theorem<br />

orbit stability, 259<br />

Bertrand’s theorem<br />

orbit solution, 252<br />

Bicycle stability<br />

roll<strong>in</strong>g wheel, 349<br />

Bifurcation<br />

non-l<strong>in</strong>ear system, 99<br />

Billiard ball, 15<br />

Bohr<br />

history, 8<br />

model of the atom, 487<br />

Bohr-Sommerfeld atom<br />

special relativity, 479<br />

Brahe<br />

history, 2<br />

Bulk modulus of elasticity, 445<br />

Buoyancy forces, 540<br />

Calculus of variations<br />

brachistochrone, 107<br />

Euler, 107<br />

history, 107<br />

Leibniz, xviii<br />

Canonical equation of motion<br />

Hamilton’s equations of motion, 198<br />

Canonical perturbation theory<br />

Hamilton-Jacobi theory, 430<br />

harmonic oscillator perturbation, 430<br />

Canonical transformations<br />

generat<strong>in</strong>g function, 410<br />

Hamilton method, 412<br />

Hamilton’s equations of motion, 409<br />

Hamilton-Jacobi theory, 414<br />

identity transformation, 412<br />

Jacobi method, 412, 414<br />

one-dimensional harmonic oscillator, 413<br />

Cartesian coord<strong>in</strong>ates, 509<br />

Cayley<br />

history, 497<br />

Center of momentum<br />

bolas, 17<br />

Center of percussion, 35<br />

Central forces<br />

two-body forces, 245<br />

Centre of mass<br />

f<strong>in</strong>ite sized objects, 12<br />

Centre of momentum<br />

relativistic k<strong>in</strong>ematics, 465<br />

Centrifugal force<br />

parabolic mirror, 293<br />

Chaos<br />

Lyapunov exponent, 99<br />

onset of chaos for non-l<strong>in</strong>ear system, 97<br />

Characteristic function<br />

Hamilton-Jacobi theory, 416<br />

Chasles’ theorem<br />

555


556 INDEX<br />

rigid-body rotation, 310<br />

Collective synchronization<br />

coupled oscillators, 389<br />

Commutation relation<br />

Poisson bracket, 399, 489<br />

Conjugate momentum, 176, 191<br />

Conservation laws<br />

angular momentum, 11<br />

l<strong>in</strong>ear momentum, 12<br />

Conservation laws <strong>in</strong> mechanics, 21<br />

Conservation of l<strong>in</strong>ear momentum<br />

explod<strong>in</strong>g cannon shell, 15<br />

many-body systems, 14<br />

Conservation of momentum<br />

billiard ball collisions, 15<br />

Conservative forces<br />

central force, 20<br />

central two-body forces, 245<br />

path <strong>in</strong>dependence, 18<br />

potential energy, 18<br />

time <strong>in</strong>dependence, 18<br />

Constra<strong>in</strong>ed motion<br />

Euler’s equations, 118<br />

geodesic motion, 126, 481<br />

Constra<strong>in</strong>t forces<br />

scleronomic, 181, 182<br />

Constra<strong>in</strong>ts<br />

k<strong>in</strong>ematic equations of constra<strong>in</strong>t, 118<br />

non-holonomic, 119<br />

partial holonomic, 120<br />

rheonomic, 120<br />

roll<strong>in</strong>g wheel, 119<br />

scleronomic, 120, 139<br />

Cont<strong>in</strong>uity equation<br />

fluid mechanics, 449<br />

Cont<strong>in</strong>uous l<strong>in</strong>ear cha<strong>in</strong>, 439<br />

Contravariant tensor, 526<br />

four vector, 468<br />

Coord<strong>in</strong>ate systems<br />

cartesian coord<strong>in</strong>ates, 509<br />

curvil<strong>in</strong>ear, 509<br />

cyl<strong>in</strong>drical basis vectors, 512<br />

cyl<strong>in</strong>drical coord<strong>in</strong>ates, 512<br />

polar coord<strong>in</strong>ates, 510<br />

spherical basis vectors, 512<br />

spherical coord<strong>in</strong>ates, 512<br />

Coord<strong>in</strong>ate transformation<br />

rotational, 515<br />

translational, 515<br />

Copernicus<br />

history, 2<br />

Correspondence pr<strong>in</strong>ciple, 481<br />

Bohr, 400, 490, 493<br />

Dirac, 400, 490<br />

Coulomb excitation, 273<br />

Coupled l<strong>in</strong>ear oscillators<br />

benzene r<strong>in</strong>g, 381<br />

cont<strong>in</strong>uous lattice cha<strong>in</strong>, 439<br />

discrete lattice cha<strong>in</strong>, 382<br />

equations of motion, 365<br />

general analytic theory, 363<br />

grand piano, 362<br />

k<strong>in</strong>etic energy tensor, 363<br />

l<strong>in</strong>ear triatomic molecule, 379<br />

normal coord<strong>in</strong>ates, 367<br />

potential energy tensor, 364<br />

superposition, 366<br />

three bodies coupled by six spr<strong>in</strong>gs, 378<br />

three fully-coupled plane pendula, 374<br />

three nearest-neighbour coupled plane pendula,<br />

376<br />

two l<strong>in</strong>early-damped coupled oscillators, 388<br />

two parallel-coupled plane pendula, 371<br />

two series-coupled oscillators, three spr<strong>in</strong>gs, 368<br />

two series-coupled oscillators, two spr<strong>in</strong>gs, 370<br />

two-series coupled plane pendula, 373<br />

viscously-damped coupled osciilators, 388<br />

Coupled oscillators<br />

collective synchronization, 389<br />

Kuramoto model, 389<br />

Covariant tensor, 526<br />

four vector, 468<br />

Cut-off frequency, 387<br />

Cyclic coord<strong>in</strong>ates<br />

conservation of momentum, 180<br />

Cyl<strong>in</strong>drical coord<strong>in</strong>ates<br />

Hamiltonian, 199<br />

d’Alembert<br />

history, 5<br />

virtual work, 107<br />

d’Alembert’s pr<strong>in</strong>ciple<br />

Lagrange equations, 135, 196, 532<br />

virtual work, 134<br />

da V<strong>in</strong>ci, Leonardo<br />

history, 2<br />

Damped l<strong>in</strong>ear oscillator<br />

response to arbitrary periodic force, 70<br />

Damped oscillator<br />

attractor, 93<br />

critically damped, 58<br />

energy dissipation, 59<br />

Fourier transform, 68<br />

free l<strong>in</strong>early damped, 56<br />

overdamped, 58<br />

Q factor, 59<br />

underdamped, 57<br />

de Broglie<br />

history, 8<br />

de Broglie waves


INDEX 557<br />

group velocity, 488<br />

Heisenberg’s uncerta<strong>in</strong>ty pr<strong>in</strong>ciple, 78<br />

matter waves, 488<br />

Delta-function analysis, 550<br />

Green’s method, 550<br />

Descartes<br />

history, 3<br />

Differential orbit equation, 250<br />

Dirac<br />

history, 8<br />

Lagrangian approach to quantum mechanics, 492<br />

Poisson brackets <strong>in</strong> quantum physics, 400, 489<br />

relativistic quantum theory, 492<br />

Discrete lattice cha<strong>in</strong><br />

cut-off frequency, 387<br />

dispersion, 387<br />

longitud<strong>in</strong>al modes, 382<br />

normal modes, 383<br />

transverse modes, 383<br />

Discrete-function analysis<br />

l<strong>in</strong>ear systems, 548<br />

Driven damped oscillator<br />

absorptive amplitude, 64, 81<br />

arbitrary periodic harmonic force, 69<br />

elastic amplitude, 64, 81<br />

energy absorption, 63<br />

Green’s method, 550<br />

harmonically driven, 60<br />

Lorentzian (Breit-Wigner) l<strong>in</strong>e shape, 65<br />

phase shift, 61<br />

resonance, 63<br />

Steady state response to harmonic drive, 61<br />

transient response, 60<br />

uncerta<strong>in</strong>ty pr<strong>in</strong>ciple, 65<br />

Eccentricity vector<br />

hidden symmetry, 259<br />

Poisson Brackets, 406<br />

two-body motion, 257<br />

E<strong>in</strong>ste<strong>in</strong><br />

General theory of relativity, 457, 480<br />

history, 7<br />

photoelectric effect, 486<br />

postulates of special relativity, 459<br />

Special theory of relativity, 457<br />

theory of relativity, 10<br />

E<strong>in</strong>ste<strong>in</strong>’s equivalence pr<strong>in</strong>ciple<br />

general theory of relativity, 480<br />

Elasticity<br />

modulus of elasticity, 51<br />

spr<strong>in</strong>g constant, 446<br />

stra<strong>in</strong> tensor, 444<br />

stress tensor, 444<br />

Electromagnetic fields<br />

field equations, 543<br />

Equations of motion<br />

analytic solution, 37<br />

successive approximation, 37<br />

Equivalence pr<strong>in</strong>ciple<br />

weak equivalence pr<strong>in</strong>ciple, 480<br />

Equivalent Lagrangians<br />

gauge <strong>in</strong>variance, 229<br />

Euler<br />

calculus of variations, 107, see Euler’s equation<br />

history, 5<br />

Euler angles<br />

def<strong>in</strong>ition, 325<br />

l<strong>in</strong>e of nodes, 326<br />

Euler’s equation of motion<br />

rigid-body rotation, 330<br />

Euler’s equations<br />

brachistrochrone, 110<br />

calculus of variations, 108<br />

catenary, 125<br />

classical mechanics, 127<br />

constra<strong>in</strong>ed motion, 118<br />

Dido problem, 125<br />

Fermat’s pr<strong>in</strong>ciple, 115<br />

generalized coord<strong>in</strong>ates, 121<br />

geodesic, 126<br />

Lagrange multipliers, 121<br />

m<strong>in</strong>imum Laplacian, 117<br />

second form, 117<br />

selection of <strong>in</strong>dependent variable, 113<br />

several <strong>in</strong>dependent variables, 115<br />

shortest distance between two po<strong>in</strong>ts, 110<br />

Euler’s equations<br />

m<strong>in</strong>imal travel cost, 112<br />

Euler’s equations of rigid-body rotation<br />

Lagrangian derivation, 331<br />

Newtonian derivation, 331<br />

Euler’s first equation, 109<br />

Euler’s hydrodynamic equation<br />

fluid dynamics, 449<br />

Faraday’s law, 542<br />

Fast light<br />

wave packets, 102<br />

Fermat<br />

history, 3, 5<br />

Fermat’s Pr<strong>in</strong>ciple, xvii<br />

Feynman<br />

history, 8<br />

Least action <strong>in</strong> quantum mechanics, 493<br />

F<strong>in</strong>ite size bodies<br />

centre of mass, 13<br />

First order <strong>in</strong>tegrals<br />

k<strong>in</strong>etic energy, 12<br />

momentum, 11<br />

Newton’s laws, 11


558 INDEX<br />

Fluid dynamics<br />

Bernoulli’s equation, 450<br />

cont<strong>in</strong>uity equation, 449<br />

Euler’s hydrodynamic equation, 449<br />

gas flow, 450<br />

ideal fluid, 449<br />

irrotational flow, 450<br />

Navier-Stokes equation, 452<br />

viscous flow, 452<br />

Fluid flow<br />

drag force, 453<br />

Navier-Stokes equation, 452<br />

Reynolds number, 453<br />

Force<br />

constra<strong>in</strong>t forces, 141<br />

generalized force, 141<br />

partition forces, 141<br />

Four vector<br />

special theory of relativity, 467<br />

Four vectors<br />

contravariant, 468<br />

covariant, 468<br />

momentum energy, 470<br />

scalar product, 468<br />

Four-dimensional space-time<br />

Riemannian geometry, 481<br />

special theory of relativity, 467<br />

Fourier analysis, 68, 545<br />

Fourier series, 545<br />

Fourier transform, 547<br />

Fourier series<br />

cos<strong>in</strong>e and s<strong>in</strong>e series, 546<br />

exponential series, 546<br />

periodic systems, 545<br />

Fourier transform<br />

Dirac delta function, 548<br />

gaussian wavepacket, 77<br />

l<strong>in</strong>early-damped l<strong>in</strong>ear oscillator, 68<br />

rectangular wavepacket, 77<br />

s<strong>in</strong>gle square pulse, 548<br />

wavepackets, 77<br />

Galilean <strong>in</strong>variance, 10<br />

Galileo<br />

history, 2<br />

Gauge <strong>in</strong>variance, 229<br />

Gauss<br />

history, 6<br />

General theory of relativity<br />

black holes, 482<br />

deflection of light, 482<br />

gravitational lens<strong>in</strong>g, 482<br />

gravitational time dilation and frequency shift,<br />

482<br />

gravitational waves, 482<br />

Mach’s pr<strong>in</strong>ciple, 480<br />

pr<strong>in</strong>ciple of covariance, 480<br />

rotation of the perihelion of mercury, 481<br />

Generalized coord<strong>in</strong>ates, 121, 131, 135, 138, 168<br />

m<strong>in</strong>imal set, 138<br />

Generalized energy, 182, 191, 232, 396<br />

Generalized energy theorem<br />

Hamiltonian mechanics, 183<br />

Generalized force, 135, 140, 170<br />

Generalized momentum, 176<br />

Geodesic motion, 126, 481<br />

Gilbert<br />

history, 2<br />

Gravitation, 38<br />

conservative, 39<br />

curl, 41<br />

determ<strong>in</strong>ation of field from potential, 41<br />

Gauss’s law, 43<br />

Newton’s laws, 44<br />

Poisson’s equation, 45<br />

potential, 40<br />

potential energy, 39<br />

potential theory, 41<br />

reference potential, 42<br />

superposition, 40<br />

uniform sphere of mass, 45<br />

Gravitational wave<br />

General Theory of Relativity, 482<br />

Green’s function method, 550<br />

Group velocity<br />

discrete lattice cha<strong>in</strong>, 387<br />

surface waves on deep water, 74<br />

wave packets, 71, 72, 101<br />

Hamilton<br />

history, 6, 497<br />

variational pr<strong>in</strong>ciple, 107<br />

Hamilton’s Action Pr<strong>in</strong>ciple<br />

Hamilton-Jacobi equation, 224<br />

Lagrange equations, 137, 196, 221, 532<br />

stationary action, 222<br />

Hamilton’s equations of motion<br />

canonical transformations, 409<br />

Hamilton’s pr<strong>in</strong>ciple function<br />

Hamilton-Jacobi theory, 415<br />

Hamilton-Jacobi equation<br />

Hamilton’s Action Pr<strong>in</strong>ciple, 224<br />

Hamilton-Jacobi theory, 414<br />

action variable, 427<br />

action-angle variables, 425<br />

central-force problem, 419<br />

free particle, 417<br />

Hamilton’s characteristic function, 416<br />

Hamilton’s pr<strong>in</strong>ciple function, 415<br />

Hamilton-Jacobi formulations, 416


INDEX 559<br />

Jacobi’s complete <strong>in</strong>tegral, 414<br />

L<strong>in</strong>dblad resonance, 431<br />

one-dimensional oscillator, 419<br />

Schrod<strong>in</strong>ger equation, 492<br />

separation of variables, 417<br />

uniform gravitational field, 418<br />

visual representation of characteristic function,<br />

424<br />

wave-particle duality, 424<br />

Hamiltonian<br />

central field, 200<br />

classical mechanics, 189<br />

conservation, 184<br />

cyclic coord<strong>in</strong>ates, 189<br />

cyl<strong>in</strong>drical coord<strong>in</strong>ates, 199<br />

def<strong>in</strong>ition, 442<br />

isotropic central force, 186<br />

l<strong>in</strong>ear oscillator on mov<strong>in</strong>g cart, 185<br />

spherical coord<strong>in</strong>ates, 200<br />

total energy, 184<br />

total energy conservation, 183<br />

two body motion, 251<br />

Hamiltonian mechanics<br />

characteristic function, 416<br />

comparison with Lagrangian mechanics, 433<br />

electron motion <strong>in</strong> electric and magentic fields,<br />

204<br />

equations of motion, 198<br />

extended formalism, 473, 476<br />

generalized energy, 182, 191, 232, 396<br />

generalized energy theorem, 183<br />

Hooke’s law for constra<strong>in</strong>ed motion, 203<br />

Legendre transform, 196, 532<br />

non-conservative forces, 238, 244<br />

observable <strong>in</strong>dependence, 401<br />

observable time dependence, 401<br />

observables, 400<br />

one-dimensional harmonic oscillator, 201<br />

plane pendulum, 55, 202<br />

Poisson brackets, 403<br />

spherical pendulum, 209<br />

Harmonic oscillator<br />

symmetry tensor, 262<br />

Heisenberg<br />

history, 8, 497<br />

uncerta<strong>in</strong>ty pr<strong>in</strong>ciple, 78<br />

Heisenberg matrix representation<br />

quantum mechanics, 489<br />

Hidden symmetry<br />

Laplace-Runge-Lenz vector, 259<br />

Hodograph<br />

<strong>in</strong>verse-square law, 258<br />

l<strong>in</strong>ear central force, 261<br />

two-body scatter<strong>in</strong>g, 275<br />

Holonomic constra<strong>in</strong>ts<br />

generalized forces, 140, 170<br />

geometric constra<strong>in</strong>ts, 118<br />

isoperimetric constra<strong>in</strong>ts, 119<br />

Lagrange multipliers, 122<br />

Hurricane<br />

Katr<strong>in</strong>a, 303<br />

Impact parameter<br />

two-body scatter<strong>in</strong>g, 256<br />

Impulsive force<br />

angular impulsive force, 35, 166<br />

translational impulse, 34, 166<br />

Inertia tensor<br />

about center of mass of uniform cube, 316<br />

about corner of uniform solid cube, 317<br />

characteristic (secular) equation, 314<br />

components, 312<br />

diagonalization, 314<br />

general properties, 319<br />

hula hoop, 319<br />

moments of <strong>in</strong>ertia, 312<br />

parallel-axis theorem, 315<br />

perpendicular-axis theorem, 318<br />

plane lam<strong>in</strong>ae, 318<br />

pr<strong>in</strong>cipal axes, 313<br />

pr<strong>in</strong>cipal moments of <strong>in</strong>ertia, 313<br />

products of <strong>in</strong>ertia, 312<br />

th<strong>in</strong> book, 319<br />

Inertial frame, 10<br />

Galilean <strong>in</strong>variance, 457<br />

Inner product<br />

tensor algebra, 527<br />

tensors, 313<br />

Inverse variational calculus, 230<br />

Irrotational flow, 450<br />

Jacobi<br />

energy <strong>in</strong>tegral, 182<br />

history, 6<br />

Jacobi’s complete <strong>in</strong>tegral<br />

Hamilton-Jacobi theory, 414<br />

Jacobian<br />

example, 532<br />

general properties, 531<br />

transformation of differentials, 531<br />

transformation of <strong>in</strong>tegrals, 531<br />

Jacobian determ<strong>in</strong>ant, 531<br />

Kepler<br />

history, 2<br />

laws of plantary motion, 255, 280<br />

K<strong>in</strong>etic energy<br />

generalized coord<strong>in</strong>ates, 181<br />

scleronomic systems, 181, 182<br />

Kirchhoff’s rules, 65, 240<br />

Kuramoto model


560 INDEX<br />

coupled oscillators, 389<br />

Lagrange<br />

calculus of variations, 107<br />

history, 5<br />

Lagrange equations<br />

d’Alembert’s pr<strong>in</strong>ciple, 135<br />

Hamilton’s action pr<strong>in</strong>ciple, 221<br />

Hamilton’s pr<strong>in</strong>ciple, 137<br />

Lagrange multipliers, 138<br />

Lagrange equations<br />

generalized coord<strong>in</strong>ates, 138<br />

Lagrange multipliers<br />

algebraic equations of constra<strong>in</strong>t, 122<br />

Euler equations, 121<br />

<strong>in</strong>tegral equations of constra<strong>in</strong>t, 124<br />

Lagrangian<br />

def<strong>in</strong>ition, 107<br />

equivalent lagrangians, 228<br />

extended formalism, 471<br />

non standard, 230, 234<br />

relativistic free particle, 474<br />

rotat<strong>in</strong>g frame, 289<br />

special relativity, 474<br />

standard, 228<br />

state space, 198<br />

time dependent, 165<br />

Lagrangian density, 440<br />

Lagrangian mechanics<br />

Atwoods mach<strong>in</strong>e, 147<br />

block slid<strong>in</strong>g on moveable <strong>in</strong>cl<strong>in</strong>ed plane, 149<br />

body on periphery of roll<strong>in</strong>g wheel, 162<br />

central forces, 143<br />

comparison with Hamiltonian mechanics, 433<br />

comparison with Newtonian mechanics, 168<br />

cyclic coord<strong>in</strong>ates, 180<br />

disk roll<strong>in</strong>g on <strong>in</strong>cl<strong>in</strong>ed plane, 144<br />

generalized coord<strong>in</strong>ates, 121, 168<br />

holonomic constra<strong>in</strong>ts, 118, 140<br />

mass slid<strong>in</strong>g on paraboloid, 154<br />

mass slid<strong>in</strong>g on rotat<strong>in</strong>g rod, 150<br />

motion <strong>in</strong> gravitational field, 142<br />

motion of a free particle, 142<br />

non-conservative forces, 243<br />

partial holonomic systems, 157<br />

plane pendulum, 187<br />

solid sphere slid<strong>in</strong>g on hemispherical surface, 161<br />

sphere roll<strong>in</strong>g down <strong>in</strong>cl<strong>in</strong>ed plane on fritionless<br />

floor, 150<br />

spherical pendulum, 151<br />

spr<strong>in</strong>g pendulum, 152<br />

sw<strong>in</strong>g<strong>in</strong>g mass connected to a rotat<strong>in</strong>g mass, 155<br />

two connected blocks slid<strong>in</strong>g without friction, 148<br />

two connected masses slid<strong>in</strong>g on rigid rail, 156<br />

two masses slid<strong>in</strong>g on <strong>in</strong>cl<strong>in</strong>ed planes, 147<br />

unconstra<strong>in</strong>ed motion, 142<br />

velocity-dependent Lorentz force, 164<br />

yo-yo, 153<br />

Lame’s modulus of elasticity, 445<br />

Legendre transform<br />

Hamiltonian and Lagrangian mechanics, 196, 532<br />

Leibniz<br />

history, 3, 5<br />

vis viva, xviii<br />

L<strong>in</strong>ear oscillator<br />

critically damped, 58<br />

driven, 60<br />

energy dissipation, 59<br />

l<strong>in</strong>ear damp<strong>in</strong>g, 56<br />

Lissajous figures, 53<br />

overdamped, 58<br />

Q factor, 59<br />

resonance, 63<br />

Steady state response of driven oscillator, 61<br />

superposition, 52<br />

transient response of driven oscillator, 60<br />

underdamped, 57<br />

L<strong>in</strong>ear systems<br />

Fourier harmonic analysis, 68<br />

L<strong>in</strong>ear velocity-dependent dissipation, 237<br />

L<strong>in</strong>early-damped l<strong>in</strong>ear oscillator<br />

characteristic frequency, 56<br />

damp<strong>in</strong>g parameter, 56<br />

Liouville’s theorem<br />

phase space, 407<br />

Lissajous figure, 53<br />

Lorentz<br />

relativistic transformation, 459<br />

Lorentz force <strong>in</strong> electromagnetism<br />

Poisson brackets, 404<br />

Lorentz transformation<br />

M<strong>in</strong>kowski metric, 468<br />

Lyapunov exponent<br />

onset of chaos, 99<br />

Mach’s pr<strong>in</strong>ciple<br />

general theory of relativity, 480<br />

Many-body systems<br />

angular momentum, 16<br />

energy conservation, 18<br />

l<strong>in</strong>ear momentum, 14<br />

Mass<br />

gravitational, 39<br />

<strong>in</strong>ertial, 38<br />

Matrix algebra, 497<br />

addition, 498<br />

adjo<strong>in</strong>t matrix, 499<br />

diagonalization, 503<br />

example of eigenvectors, 504<br />

Hermitian matrix, 499


INDEX 561<br />

history, 497<br />

identity matrix, 499<br />

<strong>in</strong>verse matrix, 499<br />

matrix multiplication, 498<br />

orthogonal matrix, 499<br />

scalar multiplication, 498<br />

secular determ<strong>in</strong>ant, 503<br />

transpose matrix, 499<br />

unitary matrix, 500<br />

Maupertuis<br />

action pr<strong>in</strong>ciple, 224<br />

history, 5<br />

Max Born, 8<br />

history, 497<br />

quantum mechanics, 491<br />

Maxwell stress tensor, 447<br />

Maxwell’s equations<br />

Gauss’s law and flux, 539<br />

Michelson and Morley experiment<br />

ether velocity, 458<br />

M<strong>in</strong>kowski metric, 468<br />

M<strong>in</strong>kowski space time<br />

special relativity, 469<br />

Modulus of elasticity<br />

bulk modulus, 445<br />

Lame’s modulus, 445<br />

Poisson’s ratio, 446<br />

shear modulus, 446<br />

Young’s modulus, 445<br />

Moment of <strong>in</strong>ertia<br />

th<strong>in</strong> door, 33<br />

Momentum<br />

angular momentum, 11<br />

l<strong>in</strong>ear momentum, 9, 15<br />

Multivariate calculus<br />

l<strong>in</strong>ear operators, 530<br />

partial differentiation, 529<br />

Navier-Stokes equation<br />

fluid flow, 452<br />

Newton<br />

equations of motion, 24<br />

history, 3<br />

laws of gravitation, 38<br />

laws of motion, 9<br />

Pr<strong>in</strong>cipia, xviii, 3<br />

Newton’s laws of gravitation, 44<br />

Newtonian mechanics<br />

conservative forces, 25<br />

constant force problems, 24<br />

constra<strong>in</strong>ed motion, 27<br />

diatomic molecule, 26<br />

l<strong>in</strong>ear restor<strong>in</strong>g force, 25<br />

perturbation methods, 37<br />

position-dependent forces, 26<br />

projectile motion, 29<br />

rocket problem, 30<br />

roller coaster, 27<br />

time-dependent forces, 34<br />

variable mass, 29<br />

velocity-dependent forces, 28<br />

vertical fall <strong>in</strong> graviatational field, 28<br />

Noether’s theorem<br />

Atwoods mach<strong>in</strong>e, 178<br />

conservation of angular momentum, 179<br />

conservation of l<strong>in</strong>ear momentum, 178<br />

diatomic molecule, 180<br />

history, 8<br />

<strong>in</strong>variant transformations, 177<br />

rotational <strong>in</strong>variance, 179<br />

symmetries and <strong>in</strong>variance, 175, 189<br />

symmetry <strong>in</strong> deformed nuclei, 180<br />

translational <strong>in</strong>variance, 178<br />

Non-conservative forces<br />

projectile motion, 243<br />

Rayleigh dissipation force, 237<br />

Non-holonomic systems<br />

non-conservative forces, 243<br />

velocity-dependent Lorentz force, 243<br />

Non-<strong>in</strong>ertial frames<br />

centrifugal force, 290<br />

Coriolis force, 290, 291<br />

effective forces act<strong>in</strong>g, 290<br />

effective gravitation , 299<br />

Foucault pendulum, 304<br />

free fall on earth, 301<br />

horizontal motion on the earth, 301<br />

Lagrangian and Hamiltonian, 289<br />

low-pressure systems, 302<br />

Newtonian mechanics, 288<br />

nucleon orbits <strong>in</strong> spheroidal potential well, 297<br />

pirouette, 294<br />

projectile fired vertically upwards, 301<br />

projectile motion near surface of earth, 298<br />

Rossby number, 302<br />

rotat<strong>in</strong>g frame, 286<br />

rotation plus translation, 288<br />

time derivatives for a rotat<strong>in</strong>g frame, 287<br />

trajectories for free motion on earth, 300<br />

translation, 285<br />

transverse, azimuthal, force, 290<br />

weather systems, 302<br />

Non-l<strong>in</strong>ear systems<br />

bifurcation, 88, 99<br />

driven damped plane pendulum, 93<br />

limit cycle, 89<br />

onset of chaos, 97<br />

period doubl<strong>in</strong>g, 96<br />

po<strong>in</strong>t attractor, 88<br />

sensitivity to <strong>in</strong>itial conditions, 97


562 INDEX<br />

soliton, 103<br />

turbulence <strong>in</strong> fluid flow, 452<br />

van der Pol oscillator, 90<br />

weak non-l<strong>in</strong>earity, 86<br />

Norbert Wiener<br />

quantum mechanics, 491<br />

Normal modes, 359<br />

Orbit equation<br />

differential orbit equation, 250<br />

free body motion, 250, 252<br />

Orbit stability<br />

Bertrand’s theorem, 259<br />

constant restor<strong>in</strong>g force, 267<br />

Hooke’s law restor<strong>in</strong>g force, 264<br />

<strong>in</strong>verse square law, 265<br />

two-body motion, 263<br />

Parallel-axis theorem<br />

<strong>in</strong>ertia tensor, 315<br />

Pascal<br />

history, 3<br />

Pauli exclusion pr<strong>in</strong>ciple<br />

quantum physics, 488<br />

Pendulum<br />

Foucault, 304<br />

plane, 55, 202<br />

plane pendulum, 187<br />

spherical, 151, 209<br />

spr<strong>in</strong>g pendulum, 152<br />

Permutation symbol, 506<br />

Perpendicular axis theorem<br />

<strong>in</strong>ertia tensor, 318<br />

Phase space<br />

harmonic oscillator, 54<br />

Liouville’s theorem, 407<br />

Phase velocity<br />

wave packets, 101<br />

wavepackets, 71, 72<br />

Philosophical developments, xix<br />

Photoelectric effect<br />

E<strong>in</strong>ste<strong>in</strong>, 486<br />

Millikan, 486<br />

Planck<br />

constant, 485<br />

history, 485<br />

Plane pendulum<br />

state space, 55<br />

Plato<br />

history, 1<br />

Po<strong>in</strong>care<br />

chaos, 85<br />

history, 7<br />

three-body problem, 85<br />

Po<strong>in</strong>care sections<br />

state-space plots, 100<br />

Po<strong>in</strong>care-Bendixson theorem<br />

non-l<strong>in</strong>ear systems, 89<br />

Poisson<br />

history, 6<br />

Poisson brackets<br />

angular momentum conservation, 401<br />

canonical transformation, 398<br />

commutation relation, 399, 489<br />

def<strong>in</strong>ition, 397<br />

fundamental, 397<br />

Hamilton equations of motion, 403<br />

<strong>in</strong>variance to canonical transformations, 398<br />

Lorentz force <strong>in</strong> electromagnetism, 404<br />

time dependence, 400<br />

two-dimensional oscillator, 405<br />

wave motion and uncerta<strong>in</strong>ty pr<strong>in</strong>ciple, 404<br />

Poisson’s ratio, 446<br />

Potential theory<br />

gravitation, 41<br />

Precession rate<br />

<strong>in</strong>ertially-symmetric rigid rotor, 336<br />

Pr<strong>in</strong>ciple of covariance<br />

general theory of relativity, 480<br />

Pr<strong>in</strong>ciple of equivalence<br />

weak pr<strong>in</strong>ciple, 39<br />

Pr<strong>in</strong>ciple of m<strong>in</strong>imal gravitational coupl<strong>in</strong>g, 481<br />

Q-factor<br />

damped l<strong>in</strong>ear oscillator, 59<br />

Quantum mechanics<br />

Heisenberg, 488<br />

Heisenberg’s matrix representation, 489<br />

Max Born, 491<br />

Norbert Wiener, 491<br />

Paul Dirac, 400, 489<br />

Pauli exclusion pr<strong>in</strong>ciple, 488<br />

Schrod<strong>in</strong>ger, 488<br />

Schrod<strong>in</strong>ger wave mechanics, 491<br />

Queen Dido’s problem, 125<br />

Radius of gyration, 36<br />

Rayleigh dissipation function, 237<br />

Rayleigh’s dissipation function<br />

Hamiltonian mechanics, 238, 244<br />

Ohm’s law, 240<br />

Reduced mass<br />

two-body motion, 247<br />

Refractive <strong>in</strong>dex, 102<br />

Relativistic Doppler effect<br />

special theory of relativity, 463<br />

Relativistic four vector<br />

scalar product, 468<br />

Restricted holonomic systems<br />

mass slid<strong>in</strong>g on hemispherical shell, 157<br />

sphere roll<strong>in</strong>g on a hemispherical shell, 159<br />

Reynolds number


INDEX 563<br />

fluid flow, 453<br />

lam<strong>in</strong>ar flow, 454<br />

turbulent flow, 454<br />

Rheonomic constra<strong>in</strong>t, 120<br />

Riemannian geometry, 481<br />

Rigid-body rotation<br />

about a body-fixed non-symmetry axis, 32<br />

about a body-fixed po<strong>in</strong>t, 310<br />

about a po<strong>in</strong>t, 309<br />

about body-fixed symmetry axis, 31<br />

about fixed axis, 309<br />

Androyer-Deprit variables, 333<br />

angular momentum, 321<br />

angular momentum about corner of a uniform<br />

cube, 322<br />

angular momentum of cube about centre of mass,<br />

321<br />

angular velocities <strong>in</strong> terms of the Euler angle velocities,<br />

329<br />

billiards, 33<br />

body-fixed axis, 31<br />

Chasles’ theorem, 310<br />

Euler equations for torque-free motion, 332<br />

Euler’s equations of motion, 330<br />

Hamiltonian approach, 333<br />

<strong>in</strong>ertia tensor, 312<br />

k<strong>in</strong>etic energy, 323<br />

k<strong>in</strong>etic energy <strong>in</strong> terms of Euler angular velocities,<br />

328<br />

matrix formulation, 313<br />

nutation, 327<br />

parallel-axis theorem, 315<br />

precession, 327<br />

rotat<strong>in</strong>g dumbbell, 332<br />

sp<strong>in</strong>, 327<br />

stability for torque-free motion, 340<br />

stability of a roll<strong>in</strong>g wheel, 349<br />

static and dynamic balanc<strong>in</strong>g, 350<br />

symmetric top about a fixed po<strong>in</strong>t, 343<br />

torque-free rotation of symmetric top, 333<br />

Rigid-body rotation about a po<strong>in</strong>t<br />

tippe top, 346<br />

Roll<strong>in</strong>g wheel<br />

symmetric rigid-body rotation , 347<br />

Rotation matrix, 515<br />

example, 517<br />

f<strong>in</strong>ite rotations, 518<br />

<strong>in</strong>f<strong>in</strong>itessimal rotations, 519<br />

proper and improper rotations, 519<br />

Rotational <strong>in</strong>variants<br />

scalar products, 329<br />

Rotational transformation<br />

rotation matrix, 515<br />

Routh<br />

Routhian reduction, 206<br />

Routhian reduction, 206<br />

cyclic and non-cyclic Routhians, 295<br />

<strong>in</strong>verse-square central potential, 212<br />

non-cyclic Routhian, 208<br />

rotat<strong>in</strong>g frames, 295<br />

rotation of a symmetric top about a fixed po<strong>in</strong>t,<br />

344<br />

Routhian, 206<br />

spherical pendulum, cyclic Routhian, 210<br />

spherical pendulum, non-cyclic Routhian, 211<br />

Routhian reduction<br />

cyclic Routhian, 207<br />

Rutherford scatter<strong>in</strong>g, 270<br />

cross section, 272<br />

distance of closest approach, 272<br />

impact parameter, 271<br />

Scatter<strong>in</strong>g<br />

energy transfer, 36<br />

Schrod<strong>in</strong>ger<br />

history, 8<br />

Schrod<strong>in</strong>ger equation<br />

Hamilton-Jacobi equation, 492<br />

Schrod<strong>in</strong>ger wave mechanics<br />

quantum mechanics, 491<br />

Scleronomic constra<strong>in</strong>t, 120, 139, 181, 182, 363<br />

Shear modulus of elasticity, 446<br />

Signal process<strong>in</strong>g<br />

coaxial cable, 70<br />

discrete-function analysis, 548<br />

Signal velocity<br />

wave packets, 71, 101<br />

Simultaneity<br />

Special theory of relativity, 461<br />

Slow light<br />

wave packets, 102<br />

Snell’s law, xvii<br />

Soliton<br />

non-l<strong>in</strong>ear systems, 103<br />

Soliton wave, 103<br />

Sommerfeld<br />

history, 8<br />

Sommerfeld atom<br />

quantum of action, 487<br />

Spatial <strong>in</strong>version transformation, 520<br />

Special theory of relativity<br />

Bohr-Sommerfeld atom, 479<br />

energy, 465<br />

extended Hamiltonian formalism, 473, 476<br />

Extended Lagrangian formulation, 471<br />

force, 465<br />

four-dimensional space-time, 467<br />

Lagrangian, 474<br />

Lorentz spatial contraction, 461<br />

M<strong>in</strong>kowski space, 469


564 INDEX<br />

momentum transformations, 464<br />

momentum-energy four vector, 470<br />

relativistic Doppler effect, 463<br />

simultaneity, 461<br />

time dilation, 460<br />

tw<strong>in</strong> paradox, 463<br />

velocity transformations, 464<br />

Spherical coord<strong>in</strong>ates<br />

Hamiltonian, 200<br />

Spherical harmonic oscillator<br />

two-body force, 259<br />

Spherical pendulum<br />

Hamiltonian mechanics, 209<br />

Lagrangian mechanics, 151<br />

Spr<strong>in</strong>g constant, 446<br />

Standard Lagrangian, 228<br />

State space<br />

Lagrangian mechanics, 198<br />

plane pendulum, 55<br />

State-space orbits<br />

Po<strong>in</strong>care sections, 100<br />

Stern Gerlach<br />

space quantization, 488<br />

Stra<strong>in</strong> tensor<br />

elasticity, 444<br />

Stress tensor<br />

elasticity, 444<br />

Strong equivalence pr<strong>in</strong>ciple<br />

general theory of relativity, 480<br />

Superposition<br />

Fourier series, 545<br />

harmonic wave analysis, 68<br />

l<strong>in</strong>ear equation of motion, 52<br />

Symmetric top<br />

Feynman’s wobbl<strong>in</strong>g plate, 338<br />

nutation , 345<br />

oblate spheroid, 335<br />

precession, 345<br />

precession rate for torque-free symmetric top,<br />

338<br />

prolate spheroid, 335<br />

rotation about a fixed po<strong>in</strong>t, 343<br />

sp<strong>in</strong>, 346<br />

sp<strong>in</strong>n<strong>in</strong>g jack, 345<br />

torque-free rotation, 333<br />

Symmetries<br />

<strong>in</strong>variance, 189<br />

Noether’s theorem, 177<br />

Symmetry tensor<br />

anisotropic harmonic oscillator, 406<br />

isotropic harmonic oscillator, 262<br />

Poisson Brackets, 406<br />

Teleology, 5, 224<br />

Tennis racket rotation<br />

asymmetric-rotor rotation, 341<br />

Tensor algebra<br />

contravariant tensor, 526<br />

covariant tensor, 526<br />

<strong>in</strong>ner product, 313, 523, 527<br />

outer product, 524<br />

transformation properties, 528<br />

Three-body problem<br />

Lagrange po<strong>in</strong>ts, 268<br />

planar approximation, 268<br />

restricted 3-body problem, 268<br />

Time dependent force<br />

nonautonomous systems, 165<br />

Time <strong>in</strong>variance<br />

conservation of energy, 182<br />

Time reversal transformation, 521<br />

Tippe top<br />

symmetric rigid-body rotation about a po<strong>in</strong>t, 346<br />

Tornadoes<br />

weather systems, 303<br />

Torque free rotation of asymmetric body, 342<br />

Total mechanical energy, 19<br />

Transformation properties of common observables, 528<br />

Translational <strong>in</strong>variance<br />

Noether’s theorem, 178<br />

Tumbl<strong>in</strong>g of an asymmetric rotor<br />

rigid-body rotation, 351<br />

Turbulence <strong>in</strong> fluid flow<br />

non-l<strong>in</strong>ear system, 452<br />

Tw<strong>in</strong> paradox<br />

special theory of relativity, 463<br />

Two-body central forces<br />

conservative forces, 245<br />

Two-body k<strong>in</strong>ematics, 274<br />

angle transformation, 276<br />

recoil energies, 278<br />

velocity transformation, 275<br />

Two-body motion<br />

angular momentum, 247<br />

apocenter, 255<br />

barycenter, 247<br />

bound orbits, 254<br />

equations of motion, 249<br />

equivalent one-body representation, 246<br />

Hamiltonian, 251<br />

<strong>in</strong>verse cubic central force, 266<br />

<strong>in</strong>verse square law, 253<br />

isotropic harmonic oscillator, 259<br />

Kepler’s laws, 255, 280<br />

Laplace-Runge-Lenz vector, 257<br />

orbit solutions , 252<br />

orbit stability, 263<br />

pericenter, 254<br />

properties of objects <strong>in</strong> solar system, 256<br />

reduced mass, 247


INDEX 565<br />

unbound orbits, 256<br />

Two-body scatter<strong>in</strong>g<br />

differential cross section, 270<br />

impact parameter, 256<br />

Rutherford scatter<strong>in</strong>g, 270<br />

total cross section, 269<br />

Two-coupled harmonic oscillators<br />

centre-of-mass oscillations, 360<br />

eigenfrequencies, 358<br />

grand piano, 362<br />

normal modes, 357, 359<br />

symmetric and antisymmetric normal modes, 360<br />

weak coupl<strong>in</strong>g, 361<br />

Uncerta<strong>in</strong>ty pr<strong>in</strong>ciple<br />

Heisenberg, 78<br />

quantum baseball, 80<br />

Uncerta<strong>in</strong>ty pr<strong>in</strong>ciple for wave motion, 78<br />

Unity of classical and quantum mechanics, 496<br />

van der Pol oscillator<br />

attractor, 90<br />

strong non-l<strong>in</strong>earity, 92<br />

weak non-l<strong>in</strong>earity, 91<br />

<strong>Variational</strong> pr<strong>in</strong>ciples<br />

calculus of variations, 107<br />

philosophy, 107, 496<br />

pr<strong>in</strong>ciple of economy, 107, 496<br />

Vector algebra<br />

l<strong>in</strong>ear operations, 505<br />

Vector differential calculus<br />

scalar differential operator, 533<br />

scalar differential operators, 533<br />

Vector differential operators<br />

curl, 536<br />

curvil<strong>in</strong>ear coord<strong>in</strong>ates, 535<br />

divergence, 536<br />

gradient, 534, 535<br />

Laplacian, 535, 536<br />

scalar product, 534<br />

vector product, 534<br />

Vector <strong>in</strong>tegral calculus<br />

curl, 541<br />

curl <strong>in</strong> cartesian coord<strong>in</strong>ates, 541<br />

curl-free field, 543<br />

divergence <strong>in</strong> cartesian coord<strong>in</strong>ates, 538<br />

divergence theorem, 538<br />

divergence-free field, 543<br />

Gauss’s theorem, 537<br />

l<strong>in</strong>e <strong>in</strong>tegral, 537<br />

Stokes theorem, 540<br />

Vector multiplication<br />

scalar product, 505<br />

scalar triple product, 507<br />

vector product, 506<br />

vector triple product, 508<br />

Vibration isolation<br />

l<strong>in</strong>early-damped oscillator, 69<br />

Virial theorem, 22<br />

Hooke’s law, 22<br />

ideal gas law, 23<br />

<strong>in</strong>versesquarelaw,23<br />

mass of galaxies, 23<br />

Virtual work<br />

d’Alembert’s pr<strong>in</strong>ciple, 134<br />

pr<strong>in</strong>ciple, 134<br />

Wave equation, 66<br />

stationary wave solutions, 67<br />

trabell<strong>in</strong>g wave solutions, 67<br />

Wave motion<br />

discrete-function analysis, 548<br />

dispersion on discrete lattice cha<strong>in</strong>, 387<br />

electromagnetic waves <strong>in</strong> ionosphere, 75<br />

group velocity for discrete lattice cha<strong>in</strong>, 387<br />

group velocity for water waves, 74<br />

group velocity of de Broglie waves, 488<br />

plasma oscillation frequency, 76<br />

uncerta<strong>in</strong>ty pr<strong>in</strong>ciple, 78<br />

water waves break<strong>in</strong>g on a beach, 74<br />

Wave packets<br />

fast light, 102<br />

Fourier transform, 77<br />

group velocity, 71, 72, 101<br />

phase velocity, 71, 72, 101<br />

signal velocity, 71, 101<br />

slow light, 102<br />

uncerta<strong>in</strong>ty pr<strong>in</strong>ciple, 78<br />

Wave-particle duality<br />

de Broglie, 488, 491<br />

Hamilton-Jacobi theory, 424<br />

Schrod<strong>in</strong>ger, 491<br />

Weak equivalence pr<strong>in</strong>ciple<br />

general theory of relativity, 480<br />

Weather systems<br />

high-pressure systems, 304<br />

low-pressure systems, 302<br />

tornadoes, 303<br />

Work<br />

def<strong>in</strong>ition, 12, 20<br />

Young’s modulus of elasticity, 445<br />

Zeeman effect<br />

weakly-coupled normal modes, 361

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!