You are on page 1of 859

Foundations of Engineering Mechanics

A. I. Lurie
Springer-Verlag Berlin Heidelberg GmbH
A. I. Lurie

Analytical Mechanics
Translated by A. Belyaev

With 92 Figures

Springer
Series Editors:
Vladimir 1. Babitsky J. Wittenberg
Department of Mechanical Engineering Institut fur Technische Mechanik
Loughborough University Universităt Karlsruhe (TH)
LEU 3TU Loughborough, Leicestershire KaiserstraBe 12
Great Britain 76128 Karlsruhe I Germany

Author:
A. 1. Lurie t

Translator:
A. Belyaev
State Technical University
of St. Petersburg
Polytekhnicheskaya 29
195251 St. Petersburg
Russia

ISBN 978-3-642-53650-2 ISBN 978-3-540-45677-3 (eBook)


DOI 10.1007/978-3-540-45677-3

Lurie, A.I.:
Analytical Mechanics / A.I. Lurie; translated by A. Belyaev. p. cm. - Berlin; Heidelberg; New York;
Barcelona; Hong Kong; London; Milan; Paris; Tokyo: Springer, 2002
(Foundations of engineering mechanics)
Includes bibIiographical references and index.
ISBN 978-3-642-53650-2

This work is subject to copyright. AlI rights are reserved, whether the whole or part of the material is
concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation,
broadcasting, reproduction on microfilm or in other ways, and storage in data banks. Duplication of
this publication or parts thereof is permitted only under the provisions of the German Copyright Law
of September 9, 1965, in its current version, and permission for use must always be obtained from
Springer-Verlag. Violations are liable for prosecution act under German Copyright Law.

Springer-Verlag is a company in the BertelsmannSpringer pubIishing group


http://www.springer.de
© Springer-Verlag Berlin Heidelberg 2002
Originally published by Springer-Verlag Berlin Heidelberg New York 2002
Softcover reprint of the hatdcover 1st edition 2002

The use of general descriptive names, registered names, trademarks, etc. in this publication does not
imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and reguIations and therefore free for general use.

Typesetting: Camera-ready copy from authors


Cover-Design: de'bIik, Berlin
Printed on acid-free paper SPIN 10859752 62/3020!kk 543 2 1 O
Anatolii I. Lurie

This book was written by a great Rw;sian scholar and teacher, A. I. Lurie,
in the period when his talent flourished.
Anatolii Isakovich Lurie was born in 1901 in Mogilev. In 1918 he grad-
uated from a hig h school (gymnasium), and was admitted to the Faculty
of Physics and Mechanics of the SaintrPetcrsburg Polytechnic In,o.;tit utc,
named after Peter the Great , where he has been working ever since. In
1939 he was conferred the degree of Doctor of Science. He headed t he De-
partment of Theoretical Mechanics through the period from 1936 to 1941,
and from 1944 to 1977 he was the Head of the Department of Dynamics and
Strength of Machines (which was renamed as t he Department of Mechanics
and Control Processes in 1960). A.I. Lurie was a Corresponding Member
of t he USSR Academy of Sciences, Di vision of Mechanics and Control Pro-
cesses. He was a member of the Presidium of t he National Committee for
Theoretical and Applied Mechanics and a member of t he National Commit-
tee for Automatic Control. A.I. Lurie was a member of the Editorial Boards
of the renowned Russian journals" Applied Mathematics and Mechanics",
and " Mechanics o f Solids" .
His scientific activity, lasting for more than half a century, has brought
remarkable achievements. He wrote a number of magnificent books:
1. Nikolai, E.L and Lurie, A.I. Vibrations of the Frame-type Foundations.
Leningrad , Moscow, Gosstroyizdat, 1933,83 pp.
2. Loitsianskii , L.G. and Lurie, A.l. Theoretical Mechanics. In three vol-
umes. Leningrad , Moscow, GMTI, 1934.
3. Lurie, A.I. Statics of Th in-walled Elastic Shells. Moscow, Gostekhiz-
dat , 1947, 252 pp_
4

4. Lurie, A.1. Some Nonlinear Problems of the Theory of Automatic


Control. Moscow, Gostekhizdat, 1951, 216 pp.
5. Lurie, A.1. Operational Calculus and its Application to the Problems
in Mechanics. Moscow, GITTL, 1951, 432 pp.
6. Lurie, A.1. Three-dimensional Problems of the Theory of Elasticity.
Moscow, GITTL, 1955. 492 pp.
7. Loitsianskii, L.G. and Lurie, A.1. A Course in Theoretical Mechanics.
In two volumes (5th edition). Moscow, GITTL, 1955, 380 pp., 596 pp.
8. Lurie, A I. Analytical Mechanics. Moscow, Nauka, 1961,824 pp.
9. Lurie, A.I. Theory of Elasticity. Moscow, Nauka, 1970, 940 pp.
10. Lurie, A.I. Nonlinear Theory of Elasticity. Moscow, Nauka, 1980,
512 pp.
The last book was written when A.I. Lurie was already seriously ill. He
did not live to see the proofs, nor did he see the original Russian edition
of the book. It has been translated into English by his son K.A. Lurie and
published by North Holland Publishers in 1990.
The original style of Lurie's scientific work manifested itself already in his
early publications; this is the ability to establish strong bonds between the
achievements of classical mechanics and the needs of modern technology.
His books are unparalleled by a number of practical applications. A. I. Lurie
became an ardent promoter of the so-called direct, or invariant, vector (and
later tensor) calculus. It is now difficult to imagine that once the relations
in theoretical mechanics were expressed and written in the cumbersome
coordinate form!
The work by A.I. Lurie in the field of application of operational calculus
to the study of stability of mechanical systems with distributed parameters
brought him a great fame. This study, as well as his direct contacts with
mathematicians stimulated research in the field of distribution of the roots
of quasi-polynomials.
Probably the greatest resonance in the world scientific community was
produced by Lurie's work on the theory of absolute stability of control
systems. The very statement of the problem was pioneering, as well as the
application of the Lyapunov function method to its solution. These results
initiated an enormous flow of scientific literature.
Professor Lurie is also the author of a number of articles and books on the
theory of elasticity. He devoted the last fifteen years of his life exclusively
to those problems. The typical feature of this works was its focus towards
obtaining analytical results. He did not pay any attention to numerical
methods that became so popular nowadays.
Professor Lurie was an extraordinary person. He was an attentive and re-
spectfullistener, but his interest sharpened when a colleague demonstrated
his own scientific ideas. This feature was especially attractive for the young
researchers and lecturers who wanted his opinion. His study was always full
of visitors seeking his advice, his review of papers, or simply his support.
He worked hard through all of his life, writing books, giving lectures, re-
5

viewing papers. He disliked and even might be hostile to the idle, though
possibly talented people.
In the spring of 1979 Professor Lurie underwent a serious surgery. It
took him the whole summer to recover after it. In September he came back
from Moscow. He looked fine. He said to me (I was already acting as the
Head of his Chair): "I am going to read my favorite "Theory of Elasticity"
course". I tried to object to this, and offered to read his lectures as well
as mine, to stimulate him to relax. He reacted rather sharply and insisted
on reading his own course. However, he was able to continue only until
October. In November, he gave up saying that it was too difficult. He died
on 12 February 1980. He was 78 yeas old.
I hope that the English-speaking reader will enjoy" Analytical Mechan-
ics" by A.I. Lurie. A good and talented pen.;on can write only a good book!

Professor Vladimir A. Palmov,


Head of Lurie's Chair
Preface

According to established tradition, courses on analytical mechanics include


general equations of motion of holonomic and non-holonomic systems, vari-
ational principles, theory of canonical transformations, canonical equations
and theory of their integration (the Hamilton-Jacobi theorem), integral in-
variants, theory of last multiplier and others. The fundamental laws of
mechanics are taken for granted and are not subject to discussion.
The present book is concerned with those issues of the above listed sub-
jects which, in the author's opinion, are most closely related to engineering
problems.
Application of the methods of analytical mechanics to non-trivial prob-
lems at the very stage of constructing the equations requires detailed knowl-
edge of the issues that are normally only briefly touched upon. With this
perspective considerable attention is paid to ways of introducing the gener-
alised coordinates, the theory of finite rotation, methods of calculating the
kinetic energy, the energy of accelerations, the potential energy of forces
of various nature, and the resisting forces. These introductory chapters,
which have to some extent independent significance, are followed by those
on methods of constructing differential equations of motion for holonomic
and non-holonomic systems in various forms. In these chapters the issues of
their interrelations, determination of the constraint forces and some prob-
lems of analytical statics are discussed as well. It is thought useful to include
geometric considerations of the motion of a material system as motion of
the representative point in Riemannian space. Further this approach is ap-
plied to the problems of perturbation theory. A special chapter is devoted to
the dynamics of relative motion illustrated by numerous applied problems.
8

This is followed by the study of canonical equations, canonical transfor-


mation and the prohlem of integration. The last chapter deals with the
Hamilton-Ostrogradsky principle, the principle of least action by Lagrange
and the theory of the perturbation of trajectories.
General methods are explained for particular examples, some of which
are not devoid of interest in our opinion. These examples include the prob-
lem of motion of a rigid body on a moving base, motion of a rigid body
with a cavity filled by fluid, the problem of rocket motion, application of
the Hamilton-Ostrogradsky principle to systems with distributed mass and
many others. Special attention is given to problems associated with the
perturbed motion of Earth satellites.
The examples analysed in the present book confirm the significance of
the methods of analytical mechanics for a wide range of applications which
was one of the primary aims of the author. In considering the examples
attention is paid to the statement of the problem and construction of the
equations of motion whilst their integration and analysis of results occupy
less space.
To facilitate reading, the book is provided with appendices to which
the reader is referred for the basic notion from matrix theory and tensor
analysis.
Equation numbering is as follows. The first number in parentheses indi-
cates the Chapter, the second - the Section, whilst the third - the equation
number in the Section. When a cross-reference is made within the same
Section only the last number is used. Both second and third numbers are
used for a cross-reference within the same Chapter. The complete number
appears when an equation from another Chapter is referred to.
The list of references contains the most important sources, a detailed list
not being the objective of the present book.
Some parts of the book are based on the lecture courses on analytical
mechanics and vibration theory taught by the author for more than twenty
years at the Faculty of Physics and Mechanics of the Leningrad Polytech-
nic Institute!. However the author hopes that students and researchers in
various fields of engineering will find this book useful.
Professor D.R. Merkin, who kindly consented to edit the book, gave a
great number of valuable suggestions to the author. A great assistance in
preparation of the manuscript and drawings was provided by A.K. Gibyan-
skaya and K.A. Lurie. It is the author's pleasure to express his deep grati-
tude to these people.

ITranslator's note: now the State Technical University of St. Petersburg


Translator's preface

The book" Analytical Mechanics" by A.1. Lurie was printed in Russian with
the edition of eighteen thousand copies and became a bibliographic rarity
within a few months. In Russia, this monograph is deservedly considered
as a classical book in mechanics. Translation of this book is a great honour
for me. Being a member of Lurie's Chair and one of his numerous pupils
I consider this activity as a debt of honour to perpetuate his memory in
mechanics. Also from a professional perspective, the translation was a very
interesting and cognitive experience.
While translating the book into English I tried to keep the author's
nomenclature which does not always coincide with that adopted in Western
books. For example, what is referred to as Hamilton's principle in the West-
ern literature on mechanics, the author calls the Hamilton-Ostrogradsky
principle for the reason explained in Section 12.2.
I am thankful to my son Nikita and my wife Olga, both of the State
Technical University of St. Petersburg, for the considerable technical and
linguistic support they gave during the translation.
I appreciate the kindness of my colleagues, Prof. B.A. Smolnikov and
Prof. Yu.G. Ispolov, who provided me with useful and profound suggestions
on the manuscript.
Finally, I would like to express my sincere gratitude to Dr. Stewart
McWilliam, from the University of Nottingham, UK who took the trou-
ble of editing the manuscript which I translated into English. I am greatly
obliged to him for his thorough correction of the galley-proofs.
Contents

Anatolii I. Lurie 3

Preface 7

Translator's preface 9

1 Basic definitions 19
1.1 Constraints . 19
1.2 Generalised coordinates . . . . . . . . . 21
1.3 Generalised velocities and accelerations 25
1.4 Redundant coordinates . . . . . . . . . 28
1.5 Quasi-velocities and quasi-coordinates . 30
1.6 Virtual displacements . . . . . . . . . . 34
1. 7 On the commutative operations of differentiation and variation 36
1.8 Variations of quasi-coordinates . . . . . . . . . . . . . . 39
1.9 Some properties of three-index symbols . . . . . . . . . 40
1.10 Calculation of three-index symbols for a two-axle trolley 42

2 Rigid body kinematics - basic knowledge 47


2.1 Rigid body position . . . . . . 47
2.2 Transformation of coordinates . 50
2.3 Euler's angles . . . . . . . . . . 51
2.4 Airplane angles and ship angles 54
12 Contents

2.5 Using matrix multiplication to obtain tables of direction


cosInes . . . . . . . . . . . . . . . . . 60
2.6 Application to Cardan's suspension. 62
2.6.1 Cardan's suspension . . . . . 62
2.6.2 Double Cardan suspension. . 64
2.6.3 The platform on a Cardan suspension 67
2.7 The velocity of a point in a rigid body . . . . 71
2.8 Vector of infinitesimal rotation . . . . . . . . 72
2.9 Angular velocity vector in terms of the time derivative of
Euler's angles . . . . . . . . . . . . . . 74
2.10 Calculation of three-index symbols . . 76
2.10.1 Sphere rolling on a rough plane 79
2.10.2 A ring rolling on a plane. . . . 79
2.11 Acceleration of a point in a rigid body 81
2.12 Matrix form for velocity and acceleration in a rigid body. 83
2.13 Differentiation of vector in a moving coordinate system 87
2.14 Relative motion. . . . . . . . . . . . . . . . . . . . . . .. 88
2.15 Absolute acceleration of point moving over the rotating earth 91
2.16 Body rolling on a fixed plane . . . . . . . . . . . . 93
2.17 Composition of motions of a rigid body . . . . . . 102
2.18 Motion of the natural trihedron of a spatial curve. 105

3 Theory of finite rotations of rigid bodies 111


3.1 Rodrigues formula and the vector of finite rotation 111
3.2 Parameters of Rodrigues and Hamilton. 114
3.3 Composition of finite rotations 117
3.4 Subtraction of finite rotations . . . . . . 122
3.5 Commutative finite rotations . . . . . . 122
3.6 Finite rotation and in terms of Euler's angles 124
3.7 Applications of formula for finite rotation . . 125
3.7.1 Rotor in Cardan's suspension. . . . . 125
3.7.2 Rotation of the geocentric system of axes 127
3.7.3 Orientation of the axis of a balanced rotor in Car-
dan's suspension relative to the geocentric system of
axes . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
3.8 Expressions for the angular velocity vector in terms of finite
rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
3.9 Cayley-Klein's parameters . . . . . . . . . . . . . . . . . . . 133
3.10 Angular velocity in terms of Cayley-Klein's parameters. .. 138
3.11 Determination of a rigid body position from angular velocity 139
3.12 Darboux's equation. . . . . . . . . . . . . . . . . . . . 142
3.13 An example. The position of a self-excited rigid body. 146

4 Basic dynamic quantities 151


4.1 Kinetic energy of a system. 151
Contents 13

4.2 Associate expression for the kinetic energy. 155


4.3 Tensor of inertia . . . . . . . . . . . . 158
4.4 Transformation of the tensor of inertia 162
4.5 The principal axes of inertia . . . 164
4.6 Inertia ellipsoid . . . . . . . . . . . . . 167
4.7 The kinetic energy of rigid body . . . 170
4.8 Principal momentum and principal angular momentum of a
rigid body . . . . . . . . . . . . . . . . . . . . . . . . 172
4.9 The kinetic energy of a system under relative motion . 174
4.10 Energy of accelerations. . . . . . . . . . . . . . . . . . 176
4.11 Energy of accelerations of a rigid body . . . . . . . . . 180
4.12 Example calculations of the kinetic energy for multi-body
systems . . . . . . . . . . . . . . . . . . . . 183
4.12.1 A gyroscope in Cardan's suspension . . . . . . . . . 183
4.12.2 A shell carrying flywheels . . . . . . . . . . . . . . . 184
4.12.3 The kinetic energy of a body carrying an unbalanced
flywheel . . . . . . . . . . . . . . . . . . 185
4.12.4 A platform carrying gimbals with rotors . . . . 188
4.12.5 Gyrovertical. . . . . . . . . . . . . . . . . . . . 191
4.13 Examples of kinetic energy and energy of accelerations 193
4.13.1 Rolling sphere . 193
4.13.2 Rolling ring . . . 194
4.13.3 Two-axle trolley 196

5 Work and potential energy 203


5.1 Generalised forces . . . . . . . . . . . . . 203
5.2 Elementary work of forces acting on a rigid body 205
5.3 Potential energy . . . . . . . . . . . . . . . . . 208
5.4 Forces that depend linearly on the coordinates 214
5.5 Potential energy due to the force of gravity 216
5.6 The shape of the Earth . . . . . . . . . . . . . 221
5.7 Elastic forces . . . . . . . . . . . . . . . . . . . 227
5.8 Calculation of the potential energy for rod structures . 231
5.8.1 A statically determinate system. . . . . . . . . 231
5.8.2 A statically indeterminate system. . . . . . . . 234
5.9 The potential energy of a rod under bending, torsion and
compression. . . . . 238
5.9.1 Plain curve . 242
5.9.2 Helical spring 245
5.10 Power . . . . . . . . 247
5.11 The dissipation function 248
5.12 Examples of the calculation of the dissipation function 252
5.12.1 Double mathematical pendulum with a square-law
resisting force . . . 252
5.12.2 Coulomb's friction . . . . . . . . . . . . . . . . . . . 254
14 Contents

5.13 Aerodynamic resisting force . . . . . . . . . . . . . . . . . . 260

6 The fundamental equation of dynamics. Analytical statics267


6.1 Lagrange's equations of the first kind. . . . . . . . . . . . . 267
6.2 Ideal constraints . . . . . . . . . . . . . . . . . . . . . . . . 271
6.3 The fundamental equation of dynamics and Lagrange's cen-
tral equation . . . . . . . . . . . . . . . . . . . 272
6.4 Rearrangement of Lagrange's central equation. . . . . . . . 275
6.5 Equilibrium of the system of particles . . . . . . . . . . . . 278
6.6 Examples of deriving equilibrium equations and constraint
forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
6.6.1 System of three rods. . . . . . . . . . . . . . . . . . 281
6.6.2 Equilibrium of a heavy rod gliding by their ends on
a smooth surface . . . . . . . . . . . . . . . . . .. 283
6.6.3 Rod in an elliptic cup . . . . . . . . . . . . . . . . . 286
6.6.4 Equilibrium of a rigid body in a central force field . 288
6.6.5 Equilibrium of a rigid body suspended on elastic rods 291
6.6.6 A special case of a prestressed system . . . . . .. 292
6.6.7 Equilibrium in the presence of Coulomb's friction . 296

7 Lagrange's differential equations 303


7.1 Derivation of Lagrange's equations of the second kind 303
7.2 The energy integral. . . . .. . . . . 307
7.3 The structure of Lagrange's equations .. 309
7.4 Explicit form of Lagrange's equations 310
7.5 Geometric interpretation of particle motion 312
7.6 Motion of a particle on a surface 316
7.7 Examples . . 319
7.7.1 Motion of a free particle relative to a non-orthogonal
coordinate system . . 319
7.7.2 Equations of motion of a free particle relative to an
orthogonal coordinate system 321
7.7.3 Equations of motion on the surface of revolution 324
7.7.4 Motion on a developable surface 325
7.8 Geometrical interpretation of the equations of motion of the
system . 326
7.9 Example of applications of Lagrange's equations 329
7.9.1 Double mathematical pendulum in the case of a mov-
ing suspension point and a quadratic resisting force 329
7.9.2 Motion of a folded string 331
7.9.3 Gyroscope in a Cardan suspension 334
7.9.4 System of two rods. 339
7.10 Determination and elimination of constraint multipliers 341
7.11 Examples 343
7.11.1 Four-rod system .. 343
Contents 15

7.11.2 Plane motion of a heavy rigid body on a string pass-


ing through a fixed ring. . . . . . . . . . . . . . . . . 346
7.12 Generalised reaction forces of removed constraints . . . . . 349
7.13 Geometrical interpretation of the generalised constraint forces353
7.14 Application to planar systems of rods . . . . . . . . . . . 356
7.14.1 Physical pendulum. . . . . . . . . . . . . . . . .. 356
7.14.2 Generalised constraint forces in plane mechanisms 358
7.14.3 Crankshaft mechanism. 363
7.14.4 System of two rods 364
7.15 Cyclic coordinates .. . . . . . 367
7.16 The Routhian function. . . . . 370
7.17 Structure of the Routhian function 373
7.18 Examples . . . . . . . . . . . . . . 377
7.18.1 Motion of a particle in a central force field (Keplerian
motion) . . . . . . . . . . 377
7.18.2 Heavy top. . . . . . . . . 379
7.18.3 System of two heavy tops 383
7.19 Quasi-cyclic coordinates . . . . . 387

8 Other forms of differential equations of motion 391


8.1 The Euler-Lagrange differential equations 391
8.2 Examples . . . . . . . . . . . . . . . . . 395
8.2.1 Sphere rolling on a rough surface 395
8.2.2 R i n g . . . . . . . . . . . . . . . . 398
8.2.3 Two-axle trolley . . . . . . . . . 400
8.3 Rolling of a rigid body on a fixed surface. 402
8.4 The case of a body bounded by a surface of revolution 408
8.5 Appell's differential equations . . . . . . . . . . . . . . 415
8.6 Appell's equations in terms of quasi-velocities . . . . . 418
8.7 Explicit form of Appell's equations. Chaplygin's equations 421
8.8 Applications to non-holonomic systems. 425
8.8.1 Sphere...... 425
8.8.2 R i n g . . . . . . . . . . . . . . . . 425
8.8.3 Two-axle trolley . . . . . . . . . 425
8.8.4 Chaplygin's equations for the problem of a rolling
sphere. . . . . . . . . . . 426
8.8.5 Plane motion of a particle . . . . . . . . 428
8.8.6 Friction gear . . . . . . . . . . . . . . . 429
8.9 Explicit forms of the Euler-Lagrange equations 434
8.10 Equations of motion of a free rigid body 436
8.11 Equations of motion of a spinning shell . 443

9 Dynamics of relative motion 449


9.1 Differential equations of motion of a carrying body . . . . . 449
9.2 Differential equations of the relative motion of carried bodies 456
16 Contents

9.3 Relative equilibrium . . . . . . . . . . . . . . . . . . . . . . 459


9.4 Equilibrium of rotating flexible shaft . . . . . . . . . . . . . 462
9.5 A gyroscope in Cardan's suspension mounted on a moving
platform . . . . . . . . . . . . . 468
9.6 Relative motion of rigid bodies . . . . . . . . . . . . . . . . 475
9.7 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . 483
9.7.1 Equations of rotation of the rigid body with a fixed
point accounting for the rotation of the Earth . 483
9.7.2 Heavy top. . . . . . . . . . . . . . . . . . . . . . . . 484
9.7.3 Rigid body carrying rotating flywheels. . . . . . . . 486
9.7.4 Oscillations of particles attached to a moving rigid
shell . . . . . . . . . . . . . . . . . . . . . . . . . . . 490
9.8 Equations of motion of a rigid body having a cavity filled by
fluid . . . . . . . . . . . . . . . 492
9.9 Equations of motion for a solid 498
9.10 Oscillations of a rotating rod . 506
9.11 Equations of motion of a rocket 512
9.12 Gyroscopic platform . . . . . . 515

10 Canonical equations and Jacobi's theorem 523


10.1 Legendre's transformation . . . . . . . . 523
10.2 Canonical equations of motion . . . . . 526
10.3 Explicit form of the canonical equations 531
10.4 Examples . . . . . . . . . . . . . . . . . 532
10.4.1 Motion of a particle in a central force field. 532
10.4.2 Canonical equations of motion of a heavy top under
given motion of the support point . . . . . . . . . . 533
10.4.3 Canonical equations of motion of a heavy top carry-
ing a flywheel . . . . . . . . . . . . . . . . 536
10.5 The Poisson brackets and the Lagrange brackets 537
10.6 Poisson's theorem. . . . . 541
10.7 Canonical transformations . . . . . . . . . . 543
10.8 Generating functions . . . . . . . . . . . . . 546
10.9 Invariance of the canonical transformations 550
1O.lOExamples of canonical transformations 552
10.10.1 First example. . . . . . . . . . . . . 552
10.10.2 Second example. . . . . . . . . . . . 553
10.11Canonical equations of the relative motion. 554
1O.12Canonical transformation and the process of motion 557
1O.13Jacobi's theorem . . . . . . . . . . . . . . . . . . . . 560
1O.14Separability of variables in the Jacobi-Hamilton equation 566
10.14.1 Keplerian motion. . . . . . . . . . . . . . . . . .. 567
10.14.2 Keplerian motion in spherical coordinates . . . .. 568
10.14.3 Motion of a particle in the field of two attracting centres569
1O.14.4Stiickel's theorem . . . . . . . . . . . . . . . . . . . . 571
Contents 17

10.14.5 Liouville's system. 574


1O.15Keplerian motion. 576

11 Perturbation theory 585


11.1 Method of parameter variation . . . . . . . . . . . . . . . . 585
11.2 Canonical equations of perturbed motion . . . . . . . . . . 589
11.3 Motion of a particle in the gravitational field of the rotating
Earth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 590
11.4 Motion of a particle in a resistive medium . . . . . . . . . . 598
11.5 Influence of small perturbations on oscillations about the
equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . 599
11.6 Influence of misbalance on the motion of a heavy top. . 605
11.7 Rotation of an Earth satellite about its centre of inertia 611
11.8 Equations of the perturbed Keplerian motion . . . . . . 621
11.9 Perturbed motion of the centre of inertia of the Earth satellite625
l1.lOVariational equations. . . . . . . . . . . . . . 631
11.110n integration of variational equations. . . . . . 634
11.12Equations for perturbed motion of a particle 637
11.13Perturbed Keplerian motion over a circular orbit 642
11.14Equations for perturbed motion of a material system 649
11.15Systems with two degrees of freedom. 652
11.16Systems with three degrees of freedom 657
11.17Stationary unperturbed motions . . . 660
11.18Examples . . . . . . . . . . . . . . . . 661
11.18.1 Two particles attached together with a string 661
11.18.2 Stability of regular precession. . . . . . . . . 665

12 Variational principles in mechanics 669


12.1 Hamilton's action. . . . . . . . . . . . . . . . . . . . 669
12.2 The Hamilton-Ostrogradsky principle . . . . . . . . 671
12.3 On the character of extremum of Hamilton's action . 676
12.4 Application to non-holonomic systems . . . . . . . . 692
12.5 Equations of motion of distributed systems . . . . . 698
12.5.1 Vibration of a hanging chain with a mass on the end 698
12.5.2 Vibration of a rotating elastic rod . . . . . . . . . . 707
12.5.3 Vibration of a chain line. . . . . . . . . . . . . . . . 710
12.6 Approximate determination of natural frequencies and nor-
mal forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . 714
12.7 Examples of approximate calculation of natural frequencies
and forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . 723
12.7.1 Vibration of a hanging chain with a mass on the end 723
12.7.2 Vibration of a rotating elastic rod . . . . . 726
12.7.3 Oscillation of the mathematical pendulum . 727
12.8 Hamilton's principal function 729
12.9 Asynchronous variation . . . . . . . . . . . . . . . 733
18 Contents

12.10The Lagrange principle of stationary action . . . . . . . . . 735


12.1IJacohi's principle of stationary action . . . . . . . . . . . . 739
12.12Metric of the element of action and metric of the kinematic
element . . . . . . . . . . . 742
12.13Perturbation of trajectories . . . . . . . . . . . 747
12.14Examples . . . . . . . . . . . . . . . . . . . . . 753
12.14.1 Trajectories of a particle under gravity. 753
12.14.2 Motion of a particle in central force field. 756
12.14.3 Motion of a particle on a conical surface . 757
12.14.4Motion on a circle in the field of attraction of two
centres. . . . . . . . . . . . . . 760
12.15Rotation of a near vertical rigid body 761
12.16Hamilton's characteristic function. . . 765
12.16.1 Motion in a gravitational field. 770
12.16.2 Keplerian motion. . . . . . . . 772
12.170n the character of the extremum of Lagrange's action. 776

A Elements of the theory of matrices 781


A.1 Definitions. . . . . . . . 781
A.2 Operations on matrices 786
A.3 Inverse of the matrix . . 794
A.4 Matrix representation of the operations of vector calculus 801
A.5 Differentiation of a matrix . . . . . . . . . . . . . . . . .. 802

B Basics of tensor calculus 805


B.1 General non-orthogonal coordinates . . . . . . . . . . . . 805
B.2 Vectors using the non-orthogonal coordinates . . . . . .. 808
B.3 Tensors of second rank in the non-orthogonal coordinates 810
B.4 Curvilinear coordinates . . . . . . . . . . . . . . . . 811
B.5 Covariant differentiation . . . . . . . . . . . . . . . . 814
B.6 Examples of non-orthogonal curvilinear coordinates. 818
B. 7 Formulae of the theory of surfaces . . . . . . . 819
B.8 Curvature of lines on the surface . . . . . . . . 823
B.9 Covariant derivative of a vector on the surface. 825
B.10 Orthogonal curvilinear coordinates 828
B.ll Finite-dimensional Euclidean space. . . . . . . 833
B.12 Riemannian space of dimension n . . . . . . . . 836
B.13 Riemannian subspace Rn in the Euclidean space En 838
B.14 The Riemann-Christoffel tensor. . . . . . . . . . . . 845

References 849

Index 857
1
Basic definitions

1.1 Constraints
From a dynamical point of view any material system can be regarded as
a collection of material particles. The relationships between the quantities
determining the position and the velocity of the system of particles are
referred to as constraints. These relationships must hold regardless of the
initial conditions and the forces acting on the system.
An example of a system subject to constraints is a rigid body which is a
collection of material particles kept at invariable distances from each other.
The invariable distances can be thought of as being provided by massless
inextensible rods connecting the particles. A system of material particles
is denoted free in the absence of any constraint. The solar system (the
sun and the planets are deemed as particles), elastic bodies and fluids are
examples of free systems.
The position of a particle Mi of a system is determined by its coordinates
in an inertial Cartesian coordinate system Oxyz. In what follows, the unit
base vectors of the coordinate axes are assumed to be orthogonal, unless
stated otherwise. The position vector olVl; is denoted by ri, when the
subscript i is 1,2, ... N, and N is the number of particles in the system.
The simplest and most important class of constraints are holonomic con-
straints. These ensure dependences between the coordinates of the system's
points and are expressed analytically in terms of following relations

fi (Xl, yl, Zl, ... , XN, YN, ZN; t) = 0 (i = 1, ... , r), (1.1.1)
20 1. Basic definitions

which are called constraint equations 1 . The number of constraint equations


is denoted by r. It is clear that r :::; 3N, where the equality condition
corresponds to an a priori prescribed motion.
Let us consider a particle M attached to the end of an inextensible string
of length l, the second end of which is fixed at the coordinate system origin,
which is subject to the constraint

indicating that the distance between particle M and the coordinate system
origin does not exceed l. Constraints of this sort are referred to as one-
sided constraints. The condition expressing a one-sided constraint is an
inequality. In what follows we consider two-sided constraints which are
given by equations.
A system is said to be holonomic when all of the constraints acting on
it are holonomic. Non-holonomic constraints express the relations existing
between the velocities of the particles, provided that these relations are
not reducible to dependences between the coordinates. A classical example
of a system subjected to non-holonomic constraints is a rigid body which
is constrained to roll on a surface without skidding. We restrict our con-
sideration here to non-holonomic constraints which correspond to linear
projections of the particle velocities. The constraint equations are

L (aikxk + bikYk + CikZk) + gi = 0 (i = 1, ... , r'), (1.1.2)


k=l

where aik, bik and Cik depend on the coordinates of the particles and time.
Equation (2) is equivalent to the follOWing

N
L (aikdxk + bikdYk + cikdzk) + gi dt = o. (1.1.3)
k=l

When this equality is not integrable, that is, it can not be reduced to a
finite equality of the form (1), it expresses a non-holonomic constraint. If

1 Equation numbering is as follows. The first number in parenthesis indicates the


Chapter, the second - the Section, whereas the third - the equation number in the
Section. Only the last number is used when a cross-reference within the same Section
is made. Both second and third numbers are used for a cross-reference within the same
Chapter. The complete number appears when an equation from another Chapter is
referred to.
1.2 Generalised coordinates 21

the following conditions


oaik oais oaik Obis oaik oCis oaik ogi
, , , ,
oXs oXk oYs = oXk oZs oXk at oXk
obik Obis obik oCis obik ogi
, , , (1.1.4)
oYs = oYk oZs oYk at oYk
oCik oCis oCik ogi
, ,
oZs oZk at oZk
are met for all k and s from 1 to N for a fixed i, the i-th equation in (3) is
integrable, as all equalities in (4) can be satisfied by casting

(1.1.5)

where Ii is a function of coordinate and time. Under this condition eq. (3)
expresses that a perfect differential dIi is equal to zero, i.e.

(1.1.6)

This relation is an equation for a holonomic constraint. Note, that equality


(4) is a sufficient condition of integrability, and is not a necessary one. When
condition (4) is not satisfied this does not imply that the i-th constraint
equation is not integrable since an integrating factor may exist, see [90].
A holonomic constraint is named stationary or scleronomic if time t
does not appear explicitly in constraint equation (1). A non-stationary
or rheonomic constraint depends on time. A non-holonomic constraint is
scleronomic if factors aik, bik and Cik in eq. (2) do not depend explicitly
on time, and in addition to this gi = O. When gi =I=- 0 a constraint is
considered as rheonomic because t is present in eq. (3) via dt even though
factors aik, bik and Cik are not time-dependent. The expediency of the above
classification of non-holonomic constraints becomes clear if we consider a
particular case in which condition (4) is met and the equation of holonomic
constraint is integrable. Then gi will be a non-zero constant and eq. (6)
takes the form

(1.1. 7)

i.e. it describes a rheonomic holonomic constraint.

1.2 Generalised coordinates


In order to shorten the forthcoming equations let us adopt the abridged
notation

Xv = ~3v-2' Yv = ~3v-l' Zv = ~3v (1/ = 1, ... , N). (1.2.1)


22 1. Basic definitions

This ensures that the holonomic constraint equations take the form

(1.2.2)

We assume that these equations are independent, i.e. there exists no func-
tion <I> such that

(1.2.3)

This means that the rank of the Jacobian matrix (A.1.7)2

alI afr
Df a~l a~l
...................... (1.2.4)
D~ alI afr
a~3N a~3N
in the domain of the variables ~1' ... , ~3N for all time t must be equal to r,
i.e. the deficiency of matrix (4) is zero. For instance, let the Jacobian

alI afr
a~l a~l
J= ....................... (1.2.5)
alI afr
a~r a~r

be non-zero. In this case, system of equations (2) is resolved for ~1' ... , ~r'
and the latter may be expressed in terms of the remaining 3N - r variables
~r+l' ... , ~3N and time t. Thus we obtain relations of the form

(1.2.6)

in which coordinates ~r+l' ... , ~3N are independent of each other. If a sys-
tem is holonomic, i.e. there are no non-holonomic constraints, the number
of degrees of freedom of the system is n = 3N - r. The remaining coordi-
nates ~1' ... , ~r are determined in terms of the independent coordinates by
virtue of eq. (6). Given non-holonomic constraints, the number of indepen-
dent parameters determining the system configuration minus the number
of equations of non-holonomic constraints, i.e. n - r', denotes the number
of degrees of freedom.
As a rule, the above way of introducing independent coordinates is not
applicable in practice. There is no need to take solely Cartesian coordi-
nates ~r+l' ""~3N' Instead of these one can introduce any other indepen-
dent quantities Ql, ... , Qn = Q3N-r which together determine the system

2The basic definitions and operations on matrices are given in Appendix A. References
to equations from Appendix A have a capital letter A as a prefix to the equation. For
example, (A.1.2) refers to the second equation of the first Section of the Appendix A.
1.2 Generalised coordinates 23

configuration. These may be distances, angles, Gaussian coordinates of the


point on a surface, areas and even quantities having no clear geometric
interpretation. It is only significant that the quantities introduced allow
determination of independent Cartesian coordinates
~T+k = ~T+k (qI, ... , qn; t) (k = 1, ... , n) . (1.2.7)
The condition for this is a non-zero Jacobian
a~T+l a~T+l
aql aqn
#0, (1.2.8)

expressing independence of the quantities ~T+k and solvability of equation


(7) for ql, ···,qn.
Substituting eq. (7) into the right hand side of eq. (6), the Cartesian coor-
dinates of the system particles can be then expressed in terms of quantities
ql, ... , qn referred to as the generalised coordinates and time t. Returning
to the original notation we can write

(1.2.9)

or for brevity

(1.2.10)
When the constraints are stationary, one can choose the generalised co-
ordinates (7) so that time t does not appear explicitly in eq. (9). In what
follows, while speaking on the subject of stationary constraints we will as-
sume this choice has been made. Then, in the case of stationary constraints
eq. (10) is written as follows

(1.2.11)
where ri denotes the position vector of point Mi in an inertial Cartesian
coordinate system.
Removing generalised coordinates from the 3N equations in eq. (9) we
arrive at 3N - n equations for the holonomic constraints. This process is
feasible as deficiency of the Jacobian matrix
a~l a~l
D~
aql aqn
....................... (1.2.12)
Dq
a~3N a~3N
aql aqn
24 1. Basic definitions

is zero due to eq. (8).


It can not be stated that the constraints are non-stationary when time t
appears explicitly in eq. (9) since the form of these equations relates to the
choice of the generalised coordinate. If removing these coordinates from eq.
(9) leads to such relationships among Cartesian coordinates (in an inertial
system) in which time t does not appears explicitly the constraints are
stationary, otherwise they are non-stationary.
The following simple example explains the aforesaid. Let us consider a
crank OM of length r rotating in plane Oxy about a fixed point O. Let a
generalised coordinate be the angle 7jJ between the crank and a line rotating
with a constant angular velocity w about the same axes. The Cartesian
coordinates of point Mare
x = r cos (wt + 7jJ) , x = r sin (wt + 7jJ) .
The constraint equation expressing the condition of the constant length of
the crank is given by
x2 + y2 _ r2 = o.
The latter is obtained by excluding 7jJ from equations for the Cartesian
coordinates and does not contain t. This implies that the constraint is
stationary. Another substantiation is obtained by choosing a generalised
coordinate in the form <p = wt + 7jJ. In this case time t does not appear in
the equations relating Cartesian and generalised coordinates, that is, these
equations take the form of eq. (11).
The above procedure for introducing generalised coordinates may not
be repeated when solving particular problems. Usually there is no need to
obtain holonomic constraint equations. It is worth choosing the generalised
coordinates which are natural for the system under consideration. Their
number must be necessary and sufficient for determining the configuration
of the system. Relations of the form of eq. (9) are constructed, if required,
by means of geometric or other reasoning.
Analytical mechanics is concerned with particles, systems with a finite
number of free (as in celestial mechanics) or constrained particles, as well
as rigid bodies or multi-body systems. The geometric configuration of such
systems are prescribed by a finite number of generalised coordinates. The
equations and approaches developed in analytical dynamics can be gener-
alised to consider a certain class of electric and electromechanical systems
whose behaviour can be described by a finite number of geometric quantities
and charges, both being viewed as generalised coordinates of the system.
The subject of analytical mechanics does not conventionally include the
study of systems whose state can not be rigorously described by a finite
number of degrees of freedom. These are problems of continuum mechanics
which can be studied via the methods of analytical mechanics provided
that it is possible to restrict the consideration to a finite number of param-
eters. For example, an elastic bar in an equilibrium position or vibrating
1.3 Generalised velocities and accelerations 25

about this position is given by an infinite countable set3 of coefficients of


a trigonometric series for the rod deflection. However, one can describe
the behaviour of such a dynamical system by approximating the elastic
bar by a finite number of these coefficients which are taken as generalised
coordinates.

1.3 Generalised velocities and accelerations


The derivatives of generalised coordinates with respect to time are referred
to as generalised velocities and are denoted (ii, ... , tin. In general, a dot de-
notes differentiation with respect to the time. For example, given a function
in terms of generalised coordinates and time

(1.3.1)

it follows that j is given by

(1.3.2)

The latter term is a partial derivative with respect to time provided that
it appears explicitly in eq. (1).
The velocity vector Vi of particle Mi is known to be the time derivative
of the position vector ri of this particle. By virtue of eq. (2) we obtain

(1.3.3)

For stationary constraints ri can be taken in the form of eq. (2.11). In this
case the latter term in eq. (3) vanishes and vector Vi becomes a homoge-
neous linear form of the generalised coordinates tis

(1.3.4)

An important consequence of eqs. (2) and (3) is that

(1.3.5)

3 A set is named countable when its elements may be numbered by natural numbers
1,2, ...
26 1. Basic definitions

From here on we shall consider the vector

(1.3.6)

which we shall call the virtual velocity. This is the velocity, found under
the assumption that time t in eq. (2.10) is fixed. Clearly, there is no need
to introduce this quantity while studying stationary constraints.
Denoting

(1.3.7)
one can cast expression (3) for the velocity vector in the form

n+l or.
Vi = L [/-iIs.
s=1 qs
(1.3.8)

This notation reduces the expressions for the non-stationary constraints


and allows one to apply the results obtained for stationary constraints in
the case of non-stationary constraints.
The second derivative of the generalised coordinates with respect to time
are referred to as the generalised accelerations iiI, ... , Qn. The acceleration
vector Wi is obtained by differentiating the velocity vector (3) with respect
to time such that

This expression is also obtained by differentiating eq. (8)

(1.3.10)

It is easy to prove the following equalities


d ori
(1.3.11)
dt oqs
Indeed, as follows from eq. (2)
1.3 Generalised velocities and accelerations 27

which yields equalities (11). Of course, similar equalities are valid for any
function of generalised coordinates and time.
Before finishing let us express the equations for the non-holonomic con-
straints (1.2) in terms of the generalised velocities. Noticing that

. _ ~ OZk . OZk
Zk - ~ 0 qs
s=1 qs
+ ot '
we obtain

(i = 1, ... , r'). (1.3.12)

Introducing the notation

(1.3.13)

we can cast eq. (12) in the form

n
L Bisqs + Bi = 0 (i = 1, ... , r') , (1.3.14)
s=1

which is equivalent to the following form

n
L B is dqs + Bidt = O. (1.3.15)
s=1

The constraint is integrable under the following condition

OBik OBis OBi


(1.3.16)
oqs' at oqs .

In the case of a stationary non-holonomic constraint, the factors Bis do


not depend on the time explicitly, and Bi = O.
When considering particular problems, equations for non-holonomic con-
straints are written in terms of generalised velocities, that is, in the form
of eq. (14).
28 1. Basic definitions

1.4 Redundant coordinates


It is sometimes worthwhile considering the configuration of a mechanical
system in terms of the quantities ql, ... qn, ... , qn+m, whose number exceeds
the necessary number n. In this case, m of these n + m parameters are
referred to as redundant generalised coordinates. Clearly, there exist m
expressions relating the n + m quantities ql, ... , qn+m and, in general, the
time such that

Fk (ql, ... , qn+m; t) = 0 (k = 1, ... , m) . (1.4.1)

These relationships, representing generalised equations for the constraints,


must be resolved in terms of m of the n+m quantities qs, i.e. the deficiency
of the Jacobian matrix

(1.4.2)
aFm aFm
aql aqn+m
must be zero. Provided that the Jacobian is non-zero

(1.4.3)

the redundant coordinates qn+1, ... , qn+m may be expressed in terms of


ql, ... , qn and time t

(1.4.4)

Expressions for the Cartesian coordinates in terms of the generalised ones


take the form

(1.4.5)

Inserting the equations for the redundant coordinates (4) into these ex-
pressions leads to the form of (2.9), the latter often being much more
complicated than (5). For example, the configuration of the double-crank
mechanism shown in Fig. 1.1 is determined by a single angle <PI' but intro-
ducing two additional angles <P2 and <P3 makes expressions for the Cartesian
coordinates very simple

(1.4.6)
1.4 Redundant coordinates 29

.r

FIGURE 1.1.

In this case the constraint equations (1) are given by

FI = al cos <PI + a2 cos <P2 + a3 cos <P3 - d = 0, }


(1.4.7)
F2 = al sin <PI + a2 sin <P2 - a3 sin <P3 = 0,

however elimination of <P2 from eq. (6), i.e. using relationships of the form
of eq. (2.9), would lead to rather cumbersome expressions.
A rigid body having a fixed point is a more general example. It is assumed
here and throughout the rest of this book that the reader is acquainted
with the basic principles of rigid body kinematics which can be found,
for instance, in [56]. The position of this body may be prescribed by nine
direction-cosines (}:ik which are cosines of the angles between rectangular
axes fixed relative to the body and rectangular axes fixed in space. The
constraint equations are six familiar relationships, three of which express
the fact that the sum of the three squares of direction-cosines with respect
to each fixed axis is equal to unity and the other three state the conditions of
mutual orthogonality of the movable axes. Expressions for the coordinates
of any point of the body in terms of the direction-cosines (}:ik are more
compact than those in which the direction-cosines are replaced by their
expressions in terms of three independent parameters, say Euler's angles.
We mention in passing that the position of the body may also be prescribed
by four Rodrigues parameters, see eq. (3.2.9) , expressed in terms of Euler's
angles due to eq. (3.6.6), relation (3.2.7) serving as constraint equation
(1). Expressions for the rigid body coordinates expressed in terms of the
Rodrigues parameters are more symmetric than those in terms of Euler's
angles.
Consideration of redundant coordinates enables a statement of the dy-
namical problem, such that the necessity to use a Cartesian coordinate
system is obviated. The configuration is determined by n + m parameters
ql, ... , qn+m and the system is subject to m holonomic and r' non-holonomic
30 1. Basic definitions

constraints whose equations are respectively


Fk (q1, ... , qn+m; t) = 0 (k = 1, ... , m), }
n+m (1.4.8)
L aks¢.s + ak = 0 (k = 1, ... , r').
s=l
The number of degrees of freedom is equal to n - r'.

1.5 Quasi-velocities and quasi-coordinates


Let the system configuration be given by n independent parameters q1, ... , qn.
In the majority of problems some linear forms of generalised velocities
¢.1, ... , ¢.n
(1.5.1)
are favoured over the generalised velocities themselves. Factors a s1, ... a sn
in eq. (1) are functions of the generalised coordinates. Let the number of
equations in (1) be n. If their number is n' < n, then one can simply adopt
Wn'+l = ¢.n'+l, ... , Wn = ¢.n.
Quantities Ws are referred to as quasi-velocities. Examples of these are
projections W1, W2, W3 of the angular velocity w or projections of the velocity
vo of a pole 0 on rectangular axes Ox'y' z' fixed relatively to the body.
The pole coordinates Xo, Yo, zo relative to some fixed set of rectangular
axes Oxyz and Euler's angles can be taken as the generalised coordinates
q1, ... , q6 of a free rigid body. Expressions for the above quasi-velocities are
respectively as follows
W1 = -a cos cp + ¢sindsincp, }
W2 = --asincp+¢sindcoscp, (1.5.2)
W3 = (p + ¢ cos d
and
vox' = ~oCYll + ~OCY12 + ~OCY13' }
vo y ' = XOCY21 + YOCY22 + ZOCY23, (1.5.3)
VO z' = XOCY31 + YOCY32 + ZOCY33,
where CYik denote directional cosines of the angles between axes Ox'y'z' and
axes Oxyz fixed in space (for example CY23 is the cosine of the angle between
axes Oy' and 0 z ). Expressions for the directional cosines in terms of the
Euler's angles are given in Table 2 of Section 2.3. Other parameters can
readily be expressed in terms of quasi-velocities. For instance, the velocity
components of any point in the body are given by the formulae
vx' = vox' + W2Z' - W3Y', }
v y ' = vO y ' + W3X' - W1 Z', (1.5.4)
Vz' = VO z' + W1Y' - W2 X',
1.5 Quasi-velocities and quasi-coordinates 31

where x', y', and z' are coordinates of the point relative to Ox' y' z'. Clearly,
they are unchanged by the body motion. It becomes evident that expres-
sions for (4) in terms of the generalised coordinates are unwieldy. Similarly,
expressions for the most important dynamic quantities like kinetic energy
of the rigid body, components of the moment of momentum which are
simple and clear in terms of the quasi-velocities become obscure when the
quasi-velocities are replaced by their expressions in terms of the generalised
velocities.
The very problem under consideration suggests the linear forms of the
generalised velocities to be taken as quasi-velocities. Given non-holonomic
constraints, it is worthwhile considering some linear form of the generalised
velocities vanishing due to the additional quasi-velocities. In other words,
provided that these linear forms are as follows

a s11;h + ... + asnqn = 0 (8 = 1, ... , r') , (1.5.5)

then equations for the non-holonomic constraints are

Ws=O (8=1, ... ,r'), (1.5.6)

where notation (1) is used.


Let us assume that the matrix whose entries are factors ask of the form

a= (1.5.7)

is non-degenerate, that is, its determinant lal is not equal to zero. The
equations in (1) are then resolved for the generalised velocities

qr = brl W 1 + ... + brnw n (r = 1, ... , n) . (1.5.8)

The matrix

b= (1.5.9)

is the inverse of matrix a, i.e.

b = a-I, ab = ba = E, (1.5.10)

where E is the identity matrix. In other words, introducing the Kronecker


delta
8 -=I- r,
(1.5.11)
8 = r,
32 1. Basic definitions

we have
n n
l.: asmbmr = bSTJ l.: amrbsm = b sr . (1.5.12)
m=1 m=1
Relationships (1) are assumed to be non-integrable. Let us recall that the
s - th quasi-velocity is said to be integrable when the following conditions

(r, k = 1, ... , n) (1.5.13)

are met. As mentioned above, failure to satisfy these conditions does not
mean that the right hand side of the expressions for Ws is not integrable.
In parallel with (1) let us consider the linear forms of differentials of the
generalised coordinates

(1.5.14)

Quantity d7r s is termed the differential of quasi-coordinates. As relation-


ships (1) are non-integrable, quantities 7r s as functions of coordinates do not
exist. For example, let wsdt be the projection of the vector () of an infinites-
imal small rotation of the rigid body about axis Ox', however there exists
no angle whose differential is (}x' (excluding a trivial case of rotation of a
body about an axis fixed in space and the body). An introduction of purely
symbolic notation 7r s referred to as quasi-coordinates is not unreasonable
since this allows one to shorten equations and wording. We denote
d7r s 0
dt = Ws = 7r s, (1.5.15)

with zero above (in place of a dot) indicating symbolic notation instead of
differentiation of 7r s with respect to time. Provided that expressions (1) (or
(14)) are integrable we arrive at the following relationships

7rs = 7rs(q1' ... , qn) (s = 1, ... , n),


which indicates a transition to new generalised coordinates only.
Let cp(q1' ... , qn) be a function of the generalised coordinates. By virtue
of (8) and (14) expression for its perfect differential can be represented in
the form
n 8cp n 8cp n n n 8cp n 8cp
dcp = L a dqr = L
r=1 qr
a l.: brs d7r = L
r=1 qr s=1 s=1
s d7r s L a brs = l.:
r=1 qr s=1
a 7r s
d7r s·

(1.5.16)
Here we adopt a symbolic yet logical notation

8cp _ ~b 8cp
- L rs (s = 1, ... , n) (1.5.17)
87r s
r=1 8qr
1.5 Quasi-velocities and quasi-coordinates 33

for coefficients in front of d7f sin eq. (16). This can be viewed as "the partial
derivative of 'P with respect to quasi-coordinate 7f s" .
In particular

ori _ ~ ori b
-- L.t- Z (1.5.18)
07f s - Z=l oqZ s·

This enables us to represent the velocity vector Vi in terms of the quasi-


coordinates
. ~ ori 0 ori ~ ori ori
Vi = ri = L.t ~ 7fs +~ = L.t ~Ws +~. (1.5.19)
s=l u7f s ut s=l u7f s ut

This also leads to the relationships


OVi ori
(1.5.20)
oWs 07f s

analogous to eq. (3.5). The relationships that are the inverse to (17) and
(18) will be used in what follows. These are given by

o'P ~ o'P ori ~ ori


~ = L.takr~' ~ = L.takr~. (1.5.21)
uqr k=l U7fk uqr k=l u7fk

Up until this point only linear homogeneous forms of generalised veloc-


ities with coefficients depending upon generalised coordinates have been
taken as quasi-velocities. A more general case is the definition of quasi-
velocities by means of linear forms with free terms a s ,n+1
Ws = a s 1tI1 + ... + asntIn + a s ,n+1 (8 = 1, ... , n) , (1.5.22)
where the factors ask may explicitly depend on time. There is no need to
repeat the above, it suffices to make use of notation (3.7) and add (22) with
an additional line
W n +1 = tIn+1 = 1 (an +1,z = bn +1,Z) . (1.5.23)
Then we arrive at the following expressions
n+1
Ws = L asktIk (8 = 1, ... , n + 1) (1.5.24)
k=l
and the above formulae hold provided that n is replaced by n + 1.
Expressions inverse to (24) are as follows
n n+1
tIr = L brk (Wk - ak,n+1) = L brkWk (r = 1, ... , n) , (1.5.25)
k=l k=l
where
n
br ,n+1 = - L brk a k,n+l, bn + 1,k = bn +1,k. (1.5.26)
k=l
34 1. Basic definitions

1.6 Virtual displacements


Generalised coordinates ql, ... , qn are functions of time which are obtained
by integration of the differential equations of motion subject to initial con-
ditions. The following set of functions of time

ql (t) , ... , qn (t) (1.6.1)

determines the true motion of the system under consideration. Differentials


dqs of the generalised coordinate represent infinitesimal small changes in
the true motion and are proportional to the time interval dt, i.e.

(1.6.2)

While stating the general principles of mechanics it is expedient to intro-


duce infinitesimal small quantities of another nature. The set of variables
(1) defines the system configuration at a given time instant t. Motion apart,
the question arises as to what set of configurations at the given time in-
stant are admitted by the system constraints. We restrict our consideration
to configurations which are infinitely close to the true ones. Infinitesimally
small increments in the generalised coordinates, designated as bql, ... , bqn,
are referred to as their variations. The above set of configurations is deter-
mined by the following quantities

q~ = ql (t) + bql, ... , q~ = qn (t) + bqn, (1.6.3)

with variations bqn being absolutely arbitrary for holonomic systems. One
can say that constraints of the system with n degrees of freedom admit oon
configurations at any instant t.
For non-holonomic constraints given by the equation
n
L askdqk + asdt = 0 (8 = I, ... , r') , (1.6.4)
k=l
variations bqk are related by r' conditions
n

Laskbqk = 0, (1.6.5)
k=l
since one must put bt = 0 in eq. (1) for any given instant t. The constraints
of this system with n - r' degrees of freedom admit oon-r' configurations.
Let us consider a particle M given by a position vector rio The change
of ri in time interval dt is determined by the differential

(1.6.6)
1.6 Virtual displacements 35

representing an infinitesimally small displacement of particle M during


the true motion of the system. This is to be contrasted with virtual dis-
placement of particle Mi denoted by 8ri. This infinitesimally small vector
represents change in the position vector of the particle when it moves from
the actual configuration to one of infinitesimally close configurations with-
out violating the constraints. The vector 8ri is calculated at a fixed time
instant t retaining terms up to first order in variation 8qk, that is

8ri = ri (ql + 8ql, ... , qn + 8qn, t) - ri (ql, ... , qn, t) = Ln

k=l
ori
-;;-8qk.
uqk
(1.6.7)

Provided that the constraints are not time-dependent, the latter terms in
expressions (6) and equations for non-holonomic constraints (8) drop out.
Differentials dri and variations 8ri are related by the same relationships
(5) and true displacement

(1.6.8)

belongs thus to the set of virtual displacements. A comparison of expres-


sions (6) and (7) as well as (4) and (5) indicates that, in the case of non-
stationary constraints, dri does not belong to this set.
What was said about the position vector holds for any function of gener-
alised coordinates and time <p (ql, ... , qn, t). Its differential is an increment
of this function due to the motion in time interval dt

(1.6.9)

whereas its variation

(1.6.10)

is an infinitesimally small change due to transition to an infinitely close


configuration at a fixed time instant.
Differentials of quasi-coordinates are determined by relations (5.14). For
more general expressions for quasi-velocities (5.22) these relations take the
form

(1.6.11)

Replacing dqk by 8qk and setting 8t = 0 at fixed instant t we obtain the


following expressions for variations of quasi-coordinates

(1.6.12)
36 1. Basic definitions

The inverse formulae are as follows

(1.6.13)

The expression for virtual displacement in terms of variations in quasi-


coordinates is given by

(1.6.14)

where "the partial derivatives of ri with respect to quasi-coordinates" are


given by eq. (5.18).
The variations considered in this Section are referred to as synchronous
variations. They are obtained by comparing the system configurations ad-
mitted by the constraints at the same time instant. One can study a more
general case of asynchronous variations in which the true configuration is
compared to an infinitely close configuration admitted by the system con-
straints at instant t + 8t.

1.7 On the commutative operations of


differentiation and variation
In the analysis that follows variations 8qs are assumed to be differentiable
functions of time. Then, by means of differentiation of expression (6.7) for
virtual displacement with respect to time we obtain

(1.7.1)

Alternatively, varying expression (3.9) for the velocity vector yields

(1. 7.2)

The difference between these expressions is

(1. 7.3)
1. 7 On the commutative operations of differentiation and variation 37

Af.'
I

FIGURE 1.2.

or, in differential form,

(1.7.4)

Let us consider now the path of a point Mi under the true motion and
mark the positions Mi and MI of the point at time instants t and t', see Fig.
----l-
1.2. Then denoting the position vector of point Mi by ri we obtain MiMI =
dri, whereas the position vector of point MI is ri + dri. Furthermore, let
Mt and MI* be positions of the point in varied system configurations at
the same time instant. These points Mt and MI* are found by means of
varying the vector positions of corresponding points of the true path, that
IS

The position vector of point Mt is ri + brio The variational principles of


mechanics and the methods of variational calculus imply consideration of a
sequence of varied positions Mt ,MI* etc. as a new varied path Ct. Taking
into account that time instants t and t + dt correspond to the positions Mt
and MI*, respectively, and assuming that the position vectors of the points
of the varied path are differentiable yields

Mt MI* = d (ri + bri) = dri + dbri.


Now by virtue of the evident equality

we have
38 1. Basic definitions

or
d8ri - 8dri = 0 (i = 1, ... , N). (1.7.5)
Then due to (4) we obtain

(i = 1, ... , N). (1.7.6)

Projecting these vector equations onto axes and using notation (2.1) for
Cartesian coordinates we have

Lanaev (d8qs - 8dqs) = 0 (v = 1, ... ,3N). (1. 7. 7)


s=1 qs
Recalling that the deficiency of the Jacobian matrix (2.12) is equal to zero
we conclude that the system of equations (7) has no non-trivial solutions
for d8qs - 8dqs. Therefore, a consequence of eq. (5) is the equalities
d8qs - 8dqs = 0 (8 = 1, ... , n). (1. 7.8)
Inversely to the latter result, eq. (5) follows from eq. (8).
Formulae (8) or, in another notation
(8qst - 8qs = 0 (8 = 1, ... ,n), (1. 7.9)
expresses the commutative operations of differentiation and variation, which
is known as the rule "d8 = 8d}'. This rule simplifies the derivation of the-
orems in mechanics as it reduces the amount of manipulation. In Hamel's
opinion, this fact accounts for the frequent use of this rule by Lagrange in
[55].
The equality d8 = 8d is a consequence of the variation rules adopted, and
one can introduce other rules under which the above rule does not hold. As
the equations of motion do not depend upon one or another variation rule
one can a priori expect that the differences d8qs - 8dqs will drop out from
the equations. We will have many opportunities to see this proved later.
It should be noted at this point that the deduced rule was based on the
introduction of varied path, that is, on the generalisation of the concept of
virtual displacement which was not inevitably required.
Under the above rule (9) relationship (3) takes the form
(1.7.10)
Let cp (ql, ... , qn; t) be some function of generalised coordinated and time.
Having repeated the derivation performed for the position vector we obtain
the result

(1.7.11)
1.8 Variations of quasi-coordinates 39

which is analogous to (3). Applying the rule (3) yields

(1. 7.12)

1.8 Variations of quasi-coordinates


We proceed now to derive expressions for (thr s)· - t5w s in which variations of
quasi-coordinates are given by eq. (6.12) whereas quasi-velocities are given
by eq. (5.1). If it is assumed that coefficients ask do not depend explicitly
upon t, we have

and consequently

Let us replace the generalised velocities qr and variations of generalised


coordinates in the second sum by quasi-velocities, eq. (5.8), and variations
of quasi-coordinates, eq. (6.13), respectively. We obtain
n

To simplify the notation we introduce the Boltzmann three-index symbols

s
'"Ytm
= ~~
~~
(aaa k _ aaa sr ) brt bkm
S
(8, t, m = I, ... , n). (1.8.2)
k=1 r=1 qr qk

Expression (1) then takes the form


n n n
(t57f st - t5ws = L ask [(t5qkt - t5qk] +L L '"Y:mWtt57f m, (1.8.3)
k=1 t=1 m=1
40 1. Basic definitions

and the first sum vanishes when the rule" d8 = 8ff' is utilised.
While deriving the latter result we assumed that the quasi-velocities are
homogeneous linear forms of generalised coordinates and that the coeffi-
cients do not depend explicitly on time. Under the more general definition
(5.22) n should be replaced in eqs. (3) and (2) by n + 1. In addition to
this we take into account eqs. (5.23)-(5.26) and make use of the following
expressions

The result is
n+l n+l n+l
(87r s t - 8ws = L ask [(8qk)· - 8qk] +L I: 'Y:mWt87rm
k=l t=l m=l
n n n n
= I: ask [(8qkt - 8qk] +L I: 'Y:mWt87rm + L c:':rt87rm , (1.8.4)
k=l t=l m=l m=l

where, by virtue of eq. (2),

(1.8.5)

It is easy to prove that the three-index symbols in relationships in eq. (4)


may be found by formulae (2). Indeed, we have

s _ n+ln+l(8
~ ~ asr _ 8
ask) b b
'Ytm - L...J L...J 8 8 rt km
k=l r=l qk qr

= ~~ (8a sr
L...J L...J 8
_ 8a sk ) b b
a rt km +~
L...J
(8a
at
sr _ 8a s,n+l) b b
8 rt n+l,m +
k=l r=l qk qr k=l qr
~
L...J
(8a s,n+l
a
8a sk ) b b
- 7ft k,n+l km +
(8a s,n+l
8t -
aas,n+l) b
8t
b
n+l,t n+l,m·
k=l qk

Taking into account eq. (5.26) yields expression (2).

1.9 Some properties of three-index symbols


The values of the three-index symbols 'Y':rtt depend only on the chosen way
of determining the quasi-velocities in terms of the generalised velocities,
1.9 Some properties of three-index symbols 41

but not on the structure and motion of the mechanical system under con-
sideration. In particular, if the integrability conditions (5.13) are met for a
prescribed value of s, all of the three-index symbols become zero by virtue
of eq. (8.2). This is to be expected because 7r s becomes a function of the
generalised coordinates and Ws becomes its derivative with respect to time.
The matrices

(1.9.1)

are skew-symmetric. Indeed, due to (8.2)

s
rmt
= ~~
~~
(aa sk _ aa sr ) b b = ~~ (aa sr _ aa sk ) b b
a a rm kt ~~ a a km rt,
k=l r=l qr qk r=l k=l qk qr
and comparison with (8.2) yields

r':nt = -r:m (8, t, m = 1, ... , n) . (1.9.2)


In particular,
r~t=O (s,t=l, ... ,n). (1.9.3)

As the diagonal elements are equal to zero we have to find ~n (n - 1)


three-index symbols for each s. Their total number is ~n2 (n - 1). The
calculation of the symbols is cumbersome provided that eq. (8.2) is used.
In each particular case it is desirable to reproduce the process of deriving
formulae (8.3) and determine r':nt as coefficients in front of products Wtb7r m
appearing in the difference (b7r s)· - bw s. There is no need to reject the rule
db = bd since the concrete values of the three-index symbols are not bound
to this rule.
Expressions for three-index symbols can be transformed into a form
which differs from (8.2). Differentiating relationships (5.12) we obtain

and substitution into (8.2) yields

s ~~
rmt = ~~asr
(abrt abrm
abkm - -a-bkt
) (1.9.4)
k=l r=l qk qk
This expression for the three-index symbols is used for obtaining equa-
tions for differences in "second derivatives of function cp (ql, ... , qn; t) with
respect to quasi-coordinates". According to definition (5.17) we have
42 1. Basic definitions

Now taking into account eq. (5.21) we obtain

and

and we arrive at the sought-for relationship

(1.9.5)

Let us consider a particular case which is important for applications and


consider m quasi-coordinates

Ws=aslll1+ ... +asmqm (s=1, ... ,m), (1.9.6)

where the coefficients a sm depend on ql, ... , qm, but not on qm+I, ... , qn. For
the sake of symmetry we adopt

Wm+j = qm+j (j = 1, ... , n - m). (1.9.7)

In this case, all of the three-index symbols

"(~q (s,t,q=1, ... ,m) (1.9.8)

are given by expressions (8.2) in which the sum is taken from 1 to m. All
the remaining symbols (for which one of the indexes s, t, q exceeds m) are
equal to zero.

1.10 Calculation of three-index symbols for a


two-axle trolley
A two-axle trolley is schematically depicted in Fig. 1.3. Its configuration
is prescribed by eight parameters, namely the coordinates x and y of the
joint B, the angles iJ and X as well as the angles of the wheel rotation
!.pI' !.p2,!.p3 and !.p4· Provided that the wheels do not slip, the system has
six non-holonomic constraints. Two constraints express the absence of the
lateral components of velocity at points A and B, while the other four
describe the absence of velocity at points where the wheels contact the
road. Thus, the system has two degrees of freedom.
1.10 Calculation of three-index symbols for a two-axle trolley 43

FIGURE 1.3.

According to eq. (5.6) we write equations for non-holonomic constraints


in the following form

WI = 0, W2 = 0, W3 = 0, W4 = 0, W5 = 0, W6 = 0, (1.10.1)

where the expressions for the quasi-velocities are as follows

WI = -xsin ('19 + X) + ycos ('19 + X), (1.10.2)

W2 = -xsin'l9+ycos'l9-l~ , (1.10.3)

W3 = X cos '19 + y sin '19 - a~ - TI <PI' (1.10.4)

(1.10.5)

W5 = x cos ('19 + X) + ysin('19 + X) - c (~+ x) - T2<P3' (1.10.6)

(1.10.7)

Equations (2) and (3) are expressions for the components of velocity at
points B and A along the corresponding wheel axes. The components

xcos '19 + y sin '19 - a~, xcos '19 + Ysin '19 + a~


in eqs. (4) and (5) are equal to the velocity ofthe centre of wheels 1 and 2,
whereas these equations express the condition of vanishing velocity at the
points where the wheels make contact with the road. Equations (6) and (7)
play an analogous role for wheels 3 and 4, the components

xcos ('19 + X) + ysin('19 + X) - c (~+ x),


44 1. Basic definitions

x cos (iJ + X) + y sin (iJ + X) + c ( 19 + X)

being the velocities of the centre of wheels 3 and 4.


We introduce additionally quasi-velocities

W7 = x cos (iJ + X) + ysin (iJ + X), (1.10.8)

Ws =X, (1.10.9)

the first being the velocity of joint B.


By means of the notation

ql = X, q2 = y, q3 = iJ, q4 = X, q4+s = <Ps' (8 = 1,2,3,4) ,


(1.10.10)

we can represent the matrix of coefficients a sr in the following form

- sin (iJ + X) cos (iJ + X) 0 0 0 0 0 0


- sin iJ cosiJ -l 0 0 0 0 0
cosiJ sin iJ -a 0 -rl 0 0 0
cosiJ sin iJ a 0 0 -rl 0 0
a=
cos (iJ + X) sin ('l9 + X) -c -c 0 0 -r2 0
cos (iJ + X) sin (iJ + X) c c 0 0 0 -r2
cos (iJ + X) sin (iJ + X) 0 0 0 0 0 0
0 0 0 1 0 0 0 0
(1.10.11)

We obtain b = a-I by solving equations (2)-(8) for generalised coordi-


nates. This presents no problem as the system structure is simple. First,
equations (2) and (8) yield

x = -WI sin (iJ + X) + W7 cos (iJ + X), } (1.10.12)


Y = WI cos (iJ + X) + W7 sin (iJ + X) ,
then we have

xcosiJ + ysin iJ = -WI sin X + W7 cos X, }


(1.10.13)
-xsiniJ + ycosiJ = WI cos X + w7sinX.

By virtue of eq. (3) we obtain 19 and then by means of eqs. (4) and (5) we
obtain generalised coordinates <PI and <P2. Finally, having 19 we determine
<P3 and <P4 from eqs. (6), (7), (8) and (9). The result is represented by the
1.10 Calculation of three-index symbols for a two-axle trolley 45

following matrix
-sin('!9 + X) 0 0 0 0 0 cos ('!9 + X) 0
cos ('!9 + X) 0 0 0 0 0 sin ('!9 + X) 0
1 1 1 .
TCos X 0 0 0 0 TsmX 0
l
0 0 0 0 0 0 0 1
a 1
b5l 0 0 0 b57 0
b= TIL Tl
a 1
b6l 0 0 0 b67 0
TIL Tl
C c 1 C
--cosX 0 0 0 b77
T2l T2l T2 T2
C C 1 C
- l cos X 0 0 0 b87
T2 T2l T2 T2
(1.10.14)
where

b5l - -1
Tl
CX
sm + -a cos X) ,
l
b57 = TIl (cos X -Ysin X) ,
b6l --
Tl
C
1 smx- -cosX
a
l
) , b67 = :1 (cos X + Ysin X) ,

b77 -1 ( I--smx
c. ) , b87 = T12 (1 + Ysin X) .
T2 l
For equations of motion we will need the three-index symbols with sub-
scripts 7 and 8. Applying eq. (9.4) we obtain

~ ~ ({)b
~ ~
r7 b
- - k8 -
()b r8 b )
- - k7
k=l r=l {)qk {)qk
8 8
"~asr ( {)b r7 r7 {)b r7
{)'!9 b38 + a()b b48 ) = "~asra'
r=l X r=l X
because all the elements of the eighth column are constant whereas those of
the seventh column depend only on '!9 = q3 and X = q4. Also b38 = 0, b48 =
1. Hence, we have

1'78 = -asl sin ('!9 + X) + as2 cos ('!9 + X) + aS3~l cos X - aS5..!.. (sinx+
Tl

ycos X) + as6 TIl (- sin X + Ycos X) - as7 T:l cos X + as8 T:l cos X

and by means of matrix (11)

1'i8 = 1, 1'~8 = sin'!9sin('!9 + X) + cos'!9cos('!9 + X) - cos X = 0,


1'~8 = - cos '!9 sin ('!9 + X) + sin '!9cos ('!9 + X) - Ycos x+sinx+ ycos X = 0
46 1. Basic definitions

etc. Next we obtain


1
"(78 = 1
-"(87 = 1, "(~8 = 0, (8 = 2, ... ,8). (1.10.15)

We find the remaining three-index symbols from the difference (87r s t -8w s .
By virtue of (14) and (9) we have
. 1
{) = Z (WI cos X - W2 + W7 sin X) , X = W8·

Therefore

-cos({)+X) [8x(~+X) -x(8{)+8x)]-


sin ({) + X) [8y (~ + X) - Y (8{) + 8x)]

-87r7 [W8+~(WICOSX-W2+W7sinx)] +
W7 [87r8 + ~ (87rl cos X - 87r2 + 87r7 sin X)]

and so on. Thus, the non-zero three-index symbols are as follows

1 _ 1 _ COSX 1 _ 1 _ 1 1 _ 1_
"(17 - -"(71 - --Z-, /27 - -"(72 -Z' "(78 - -"(87 - 1,
2 _ 2 _ 1 2 _ 2 _ sin X 2 _ 2 _ cos X
"(71 - -"(17 - Z' "(12 - -"(21 - -Z-, "(27 - -"(72 - -Z-,
k _ k _ cos X k _ k _ sin X
"(12 - -"(21 - -Z-, /72 - -"(27 - -Z-, (k = 3,4),
k _ k _ 1 k _ k _ sin X _
"(12 - -"(21 - Z' "(71 - -/17 - -Z-, "(81
k
-
k_
-"(18 - 1, (k = 5,6,7,8) .
(1.10.16)

Other examples of the calculation of three-index symbols, as well as


examples of deriving equations for non-holonomic constraints, are given in
Chapter 2, namely in Sections 2.10 and 2.16.
2
Rigid body kinematics - basic
knowledge

2.1 Rigid body position


As mentioned in Chapter 1 rigid bodies and systems of rigid bodies are
the most important objects of analytical mechanics. For this reason it is
a worthwhile exercise to briefly review the basic formulae for rigid body
kinematics and discuss some special problems in greater detail.
Let Oxyz be an inertial (fixed) right-handed system of rectangular axes.
In addition to this let us consider a system of rectangular axes Ox' y' z'
fixed relatively to the body and moving with it. The origin (in sequel also
referred to as the pole) 0 of system Ox'y' z' is arbitrary. To simplify some
equations we will also use the notation OXIX2X3 for the inertial system
and Ox~ x~x; for the system fixed in the rigid body. When it is desirable
to avoid primes we use the notation O~'TJ( (or O~ 1 ~2~3) for inertial axes
and Oxyz (or OXIX2X3) for moving axes. The position of the system of
moving axes Ox' y' z' is given by the coordinates of its origin Xo, Yo, zo (or
-=--+
its position vector TO = 00) and the table of direction cosines

II II x y z II
II x' II au a12 a13 II
II y' II a21 a22 a23 II
II z' II a31 a32 a33 II

Table 1 of direction cosines


48 2. Rigid body kinematics - basic knowledge

The unit vectors of the axes of the system Oxyz are denoted as is (s =
1, 2, 3), whereas i~ denote the unit vectors of the system Ox' y' z'. As follows
from the above table ask is the projection of i~ on the i k axis, i.e.

ask = i~ . ik S, k = 1,2,3, (2.1.1)

where the dot indicates the scalar product.


It follows that vector i~ can be expanded in terms of the basis vectors of
the system Oxyz
3
i~ = a s 1i 1 + a s 2i 2 + a s3 i 3 = L askik' (2.1.2)
k=l

Vectors iI, i 2, i3 are unit base vectors and produce an orthogonal set. By
means of the Kronecker delta one can cast it in the form

•• >: { 01 (k # m), (2.1.3)


lk . 1m = Ukm =
(k = m).
Hence
3 3 3
i~ . i~ = LL askatlDkl = L askatk·
k=ll=l k=l

Since the trihedral i~ is also orthogonal we obtain


3

L askatk = Dst· (2.1.4)


k=l

The latter yields six relations linking the nine direction cosines of Table
1. These relations are given by

ail + aI2 +aI3 = 1, all a21 + a12a22 + a13a23 = 0, }


a§1 + a§2 + a§3 = 1, a21 a31 + a22a32 + a23a33 = 0, (2.1.5)
a~l + a~2 + a~3 = 1, a31 all + a32a12 + a33a13 = O.

The vector product i1 x i2 is defined as a unit vector perpendicular to


the plane containing i1 and i2 and points in the direction of axis Oz, in
other words i1 x i2 = i 3. The notation

(2.1.6)

etc. can be shortened by introducing the Levi-Civita symbol Eklm, the value
of which is given by

+1 if k, l, m is an even permutation of 1,2,3,


-1 if k, I, m is an odd permutation of 1,2,3,
o otherwise.
2.1 Rigid body position 49

Formulae (6) take the form

(2.1. 7)

It is clear that

The same equations can be written for the unit vectors i~, i~, i;

(2.1.8)

Expressing in the latter equation the unit base vectors of the moving
axes h, h, h by means of eq. (2) we obtain
3 3 3
L arqiq = frst L L askatlfklqiq ,
q=l k=ll=l

i.e. the nine equations


3 3

a rq = frst L L askatlfklq· (2.1.9)


k=ll=l

For example
3 3
a12 = fIst L L ask a tl f kl2 = fIst (a s 3 a tl - a s1 a t3)
k=ll=l

or

Hence, in the determinant

a11 a12 a13


lal = a21 a22 a23 (2.1.10)
a31 a32 a33

each element is equal to the corresponding cofactor. For instance, the ele-
ments of the first row in the determinant are equal to the projections of the
unit base vector i~ on the axes Oxyz, whereas the corresponding cofactors
are equal to the projections of the vector product i~ x i; on the axes Oxyz.
Therefore

Ia I = 11·
of (Of
12 X 13
Of) = 11
° °
. 11 = 1. (2.1.11)

The matrix

a = Ilaikll (i,k = 1,2,3) (2.1.12)


50 2. Rigid body kinematics - basic knowledge

is referred to as the rotation matrix. This matrix makes the Cartesian base
Oxyz coincident with the Cartesian base Ox' y' z' (both bases are assumed
to have the same origin). It follows from the above that this matrix is
orthogonal and the inverse of the rotation matrix coincides with the trans-
posed rotation matrix (see Sec. A.4)
(2.1.13)
With help of the Levi-Civita symbol one can write the vector product of
two vectors a and b in the form
3 3 3 3

a x b = LL asbris x ir = LL asbrEsrtit· (2.1.14)


s=lr=l s=lr=l

2.2 Transformation of coordinates


-=---'
Let r denote the position vector OM of a rigid body point M in an inertial
coordinate system Oxyz and x, y, z denote the projections of the position
vector on the coordinate system axes. Let r denote the position vector
oM of point M and x', y', z' denote the coordinates of point M in the
coordinate system Ox' y' z' fixed in the body, that is x', y' , z' do not change
during the body motion. Thus
(2.2.1)
-=---7
Let ro = 00 denote the position vector of the origin O. Then projecting
an obvious geometrical identity

r = ro + r' = ro + x'i~ + y'i~ + z'i~, (2.2.2)


on the axes of the inertial coordinate system we arrive at the following
equations

The scalar form of this equations is given by

x = Xo + c¥ux' + C¥21Y' + C¥31Z', }


Y = Yo + C¥12 X ' + C¥22Y' + C¥32 Z ', (2.2.3)
Z = Zo + C¥13X' + C¥23Y' + C¥33Z'.

To derive the inverse transformation, one projects relationship (2) on


the axes of system Ox' y' z'. Multiplying both sides of eq. (2) by i~, i~, i~ we
obtain
x' = C¥u (x - xo) + C¥12 (y - Yo) + C¥13 (z - zo), }
y' = C¥21 (x - xo) + C¥22 (y - Yo) + C¥23 (z - zo) , (2.2.4)
z' = C¥31 (x - xo) + C¥32 (y - Yo) + C¥33 (z - zo) .
2.3 Euler's angles 51

FIGURE 2.1.

Let ~ and x denote column matrices having elements x, y, z and x', y', z',
respectively. We have to use this notation since a prime implies transposi-
e
tion of a matrix, for example and x' are row matrices.
Formula (4) takes the form
x = 0: (~ - ~o) . (2.2.5)
Due to orthogonality of the rotation matrix 0: eq. (3) is as follows
~=~o+o:'x. (2.2.6)

2.3 Euler's angles


In the sequel we will speak about initial position of a rigid body in which
axes Ox'y'z' fixed in the body coincide with axes Oxyz fixed in the space
(both set of axes having the same origin) and the actual position. Trihedral
Ox' y' z' can be brought from its initial position to the actual position by
three successive rotations about properly chosen axes. The angles of these
rotations known as Euler's angles are three independent parameters en-
abling calculation of the nine direction cosines O:sk (Table 1) interrelated
by six equations (1.5).
Let us agree once and for all that the positive direction of a turn is
clockwise (counterclockwise) for a right-handed (left-handed) coordinate
system for an observer watching from the positive end of the rotation axis.
It is also assumed that speaking about a turn one implies a positive turn.
The first rotation is a rotation through an angle 'ljJ (0 :S 'ljJ :S 27r) about
axis Oz. The unit base vectors i and i2 remaining in plane Oxy now coincide
52 2. Rigid body kinematics - basic knowledge

with unit base vectors nand nl, see Fig. 2.1. Vector n lies along a straight
line referred to as the nodal axis ON and determines the positive direction
for this axis. This leads to the equations

n = il cos 7/J + i2 sin 7/J, nl = -h sin 7/J + i2 cos 7/J. (2.3.1)

The second rotation through an angle {j (0 :::; {j :::; 7r) about the nodal
axis brings vectors nl and h which are perpendicular to the nodal axis into
coincidence with n' and i;, the latter defining the final position of axis Oz'.
Now we have

n' = nl cos {j + h sin {j, i;= -nl sin {j + i2 cos {j. (2.3.2)

Since the plane containing vectors nand n' is perpendicular to 0 z' it


coincides with plane Ox'y'. A further rotation through an angle 'P (0 :::;
'P :::; 27r) about axis Oz' makes vectors nand n' coincide with the unit base
vectors i~ and i~ of the axes Ox' and Oy'. Thus we have

i~ = n cos 'P+n' sin 'P, i; = -n sin 'P + n' cos 'P. (2.3.3)

Angles 7/J, {j and 'P are referred as to the angles of precession, nutation and
spin. Using relationships (1), (2) and (3) it is easy to obtain expressions for
direction cosines ask of the angles between trihedrals Ox' y' z' and Oxyz.
For example,

all i~ . h = (ncos'P + n' sin'P)· il


(il cos 7/J + h sin 7/J) . il cOS'P + (nl cos {j + h sin {j) . il sin 'P
cos 7/J cos 'P + n 1 . h cos {j sin 'P = cos 7/J cos 'P - sin 7/J cos {j sin 'P

etc. Table 1 of the direction cosines, takes now the following form

x y z
cos 'P cos 7/J- cos 'P sin 7/J+
sin 'P sin 7/J cos {j sin 'P cos 7/J cos {j
- sin 'P cos 7/J- - sin 'P sin 7/J+
cos 'P sin 7/J cos {j cos 'P cos 7/J cos {j
zl sin {j sin 7/J - sin {j cos 7/J

Table 2 of the direction cosines

While determining Euler's angles it was significant to construct two or-


thonormalised vector bases: a "half-fixed" one n, nl' h and a "half-movable"
one n, n', i;. The described way of introducing the angles is certainly not
unique. There are a number of versions used in the dynamics of aircraft,
2.3 Euler's angles 53

.··n

FIGURE 2.2.

ships and gyroscopes. In order to gain some insight into them one can be
guided by the following general ideas:
a) The principal axes are suggested, the first being fixed in space and the
second being fixed in the body. They may have the same name (like 0 z and
Oz' as in the above example) or different names (e.g. Oy and Ox'). The
planes perpendicular to the principal axes are named the principal planes,
and the reference axes are chosen in this plane (planes Oxy and Ox' y' and
axes Ox and Ox' in the above example).
b) The nodal axis, i.e. a line intersecting the two principal planes, is
introduced and a unit base vector n of this axis is taken. The simplest
way to achieve this vector is to project the moving principal axis on the
fixed principal plane and to determine the unit base vector -nl along this
projection. Then n lies in the fixed principal plane and is perpendicular to
nl·
In the moving principal plane the unit base vector n', which is perpen-
dicular to n, is constructed. Now vector n' is along the projection of the
fixed principal axis on the moving principal plane.
c) The three angles determining the position of the system Ox' y' z' are:
1) the angle between the fixed and moving principal axes. As an alterna-
tive one can take the angle between vector -nl and the moving principal
plane (their sum is 7f/2);
2) the angle between the reference axis in the fixed principal plane and
vector n (sometimes n 1 or - n 1); and
3) the angle between vector n (or n') and the reference axis in the moving
principal plane.
The above are angles of rotation about n (or -n) , fixed principal axis
and moving principal axis, respectively. In addition to this, the first angle
takes values from the interval [0, 7f] (or [-7f / 2, 7f / 2]) , whereas the other two
vary from 0 to 27f.
54 2. Rigid body kinematics - basic knowledge

!I

o'~ ____--=-

FIGURE 2.3.

A well-known example of Euler's angles in astronomy uses the following


set of angles: nand i determine the position of the orbital plane whilst w
provides the direction for a reference axis in this plane, see Fig. 2.2. The
first angle n is the longitude of the ascending node N of a planet and plays
the role of the precession angle. The second angle i is the angle between
the orbital plane and the reference fixed plane O~ry and is the nutation
angle. Angle w describes the spin and if the reference axis is directed in the
pericentron II then w describes the angular distance of the pericentron to
the ascending node.

2.4 Airplane angles and ship angles


It is common practice in airplane dynamics that the fixed axes are directed
as follows: a~ along the required heading in the horizon plane, ary along the
a
vertical uprising from point and a(to the right for an observer watching
in the direction of a~. The origin is placed in the take-off position. Axis Ox
of the airplane axes Oxyz is directed along the airplane axis from the tail
to the cockpit, axis Oy lies in the airplane symmetry plane perpendicular
to Ox and finally 0 z is perpendicular to this plane to the right for the
pilot, see Fig. 2.3.
Let us place the origin a of the coordinate system a~ry( at 0 and view
Ory and Ox as principal axes. Projecting (as indicated in Sec. 2.3) Ox on
the principal plane O(~ we obtain vector -n1 and vector n, the latter lying
in the same plane, in such a way that iz, nand n1 form an orthonormalised
right-handed trihedral, see Fig. 2.4a. Vector n' lies along the projection of
iz on the principal plane Oxyz. The simplest way to obtain this is to find
2.4 Airplane angles and ship angles 55

'I

FIGURE 2.4.

the direction perpendicular to i~ in the plane of the principal axes Ox and


Ory (in passing we note that vector nl also lies in this plane).
Angles 'IjJ, {) and cp referred to as the angles of yaw, pitch and roll are
generated as follows: i) 'IjJ is the angle between O~ and -nl while rotating
about Ory, ii) rotation through an angle cp about the axis Ox makes vector
n' coincidental with Oy, and iii) the pitch angle {) is determined as the
angle between vectors -nl and i~ while rotating about-no
The velocity axes Ox*y* z* are used in airplane dynamics along with the
airplane axes. Axis Ox* is directed along the velocity vector of point 0
which is routinely the centre of inertia of the airplane, Oy* lies in the plane
of symmetry of the aircraft perpendicular to Ox*, and 0 z* is perpendicular
to the plane containing the axes Ox* and Oy* producing a right-handed
system. The position of the velocity axes with respect to the fixed axes
O~ry( is described by angles A, J-L and v which are determined exactly in
the same way as the airplane angles 'IjJ, {) and cp. Angles>. and J-L describe
the direction of the velocity vector (that is axis Ox*) are referred to as the
angles of heading and ascend whereas angle v is called the velocity roll.
Provided that the velocity axes are prescribed, one can construct the
airplane axes by introducing the angle of slide f3 and the angle of attack
Q. By rotating the plane Ox* z* through an angle f3 about axis Oy* one

determines the unit base vectors m and ml, see Fig. 2.4b, then a rotation
through an angle Q makes vector ml coincide with the airplane axis Ox.
Recalling now that the plane of axes Ox and Oy* is the plane of symmetry
for the aircraft we find axis Oy in this plane. The angle between Oy and
Oy* is Q, while the angle between Oy and ml is 7r /2 + Q.
56 2. Rigid body kinematics - basic knowledge

In order to calculate the table of direction cosines of the angles between


the airplane axes and the fixed axes we construct the auxiliary equations

-n = h cos'lj; + i l sin 'Ij;, -nl = -i3 sin 'Ij;+il cos 'Ij;, }


n'= nl sin {) + h cos{), i~= -nl cos {)+h sin {), (2.4.1)
i~ = n' cos<p - nsin<p, i~= -n' sin <p - n cos <p,

and then we find

i~ h sin {) + cos {) (-i3 sin'lj; + i l cos {))


h cos'lj; cos {)+h sin {)-i3 sin'lj; cos {)
which defines the first row of the table of direction cosines. By analogy we
obtain

i~ cos <p (nl sin {) + h cos {)) + sin <p (h cos 'Ij; + i l sin 'Ij;)
cos <p sin {) (i3 sin 'Ij; - h cos'lj;) + h cos {)cos <p +
h sin <pcos 'Ij; + h sin'lj;sin<p,
where the factors of h, i2 and i3 comprise the second row of the table etc.
Consequently we arrive at the following table

II II ~ 'TJ ( II
II x III cos'lj; cos {) sin {) - sin 'Ij; cos{) II
sin <psin 'Ij;- sin<pcos'lj;+
cos{) cos <p
cos <p cos 'Ij; sin {) cos <p sin 'Ij; sin {)
cos <psin 'Ij;+ cos <p cos'lj;-
- cos {) sin <p
sin <p cos 'Ij; sin {) sin <p sin 'Ij; sin {)

Table 3 of the direction cosines of the airplane angles

In order to obtain the table of direction cosines of the angles between


the velocity axes and the fixed axes it is sufficient to replace the angles 'Ij;, {)
and <p in Table 3 by .x, J.L and 1/. The table of direction cosines of the angles
between the velocity axes and the airplane axes is given by

II II x y I z II
II x* II cos a cosf3 - cosf3 sin a I sinf3 II
II y* II sin a cos a I 0 II
II z* II -cosasinf3 I sinf3 sin a I cosf3 I

Table 4 of the direction cosines of the velocity angles


2.4 Airplane angles and ship angles 57

By virtue of Tables 3 and 4 one can obtain relationships expressing the


angles A, J-t and v in terms of 'ljJ, '19, rp and angles a, (3. For example,

cos (x*,~) = cos,x cos J-t = cos (x*, x) cos (x,~) +


cos (x*, y) cos (y,~) + cos (x*, z) cos (z,~)

etc. These formulae are significantly simplified if the angles of attack and
slide are assumed to be small. In this case Table 4 is replaced by the
following table

II II x I y Iz II
II x* II 1 I -a I (3 II
II y* II a I 1 I 0 II
II z* II -(3 I 0 I 1 II

Table 5 of the direction cosines

By means of this table we obtain

cos,x cos J-t = cos 'ljJ cos '19 - a (sin'ljJ sin rp - cos'ljJ cos rp sin '19) +
(3 (sin'ljJ cos rp + cos'ljJ sin rp sin '19) ,

cos J-tcos v = a sin '19 + cos '19 cos rp, sinJ-t = sin '19 - cos '19 (a cos rp + (3 sin rp) .
Because the differences between ,x, J-t, v and 'ljJ, '19, rp are small values of the
order of a and (3 we have

cos,x cos J-t cos (,x - 'ljJ + 'ljJ) cos (J-t - '19 + '19)
~ [cos'ljJ - (,x - 'ljJ) sin'ljJ] [cos '19 - (J-t - '19) sin '19]
~ cos'ljJcos'l9- (,X-'ljJ)sin'ljJcos'l9- (J-t-'I9)cos'ljJsin'l9,

cos J-t cos v ~ cos '19 cos rp - (J-t - '19) sin '19 cos rp - (v - rp) cos '19 sin rp,

sin J-t i'::j sin '19 + (J-t - '19) cos '19.

Comparing the corresponding equations we arrive at the following relation-


ships

(,x - 'ljJ) sin'ljJ cos '19 + (J-t - '19) cos 'ljJ sin '19 ~
(a sin rp - (3 cos rp) sin 'ljJ - (a cos rp + (3 cos rp) cos'ljJ sin '19,
58 2. Rigid body kinematics - basic knowledge

FIGURE 2.5.

(/1 - '!9) cos cp sin '!9 + (v - cp) cos '!9 sin cp ~ -0: sin '!9,

(/1 - '!9) cos '!9 ~ - (0: cos cp + ;3 sin cp) cos '!9,

which lead to the formulae relating the velocity and airplane angles

/1 - '!9 = - (0: cos cp + ;3 sin cp) , }


1
). - 'l/J = ----:<i (0: sin cp - ;3 cos cp) , (2.4.2)
cOSu
v - cp = (-o:sincp + ;3coscp)tan'!9.

Ship axes differ from airplane axes only in notation. Figure 2.5 shows
construction of the ship axes Oxyz suggested by Krylov; axis Ox is directed
from the aft to the fore, axis Oy is directed to the port side and axis 0 z
lies in the centre plane of the ship. They coincide with the axes O~rt( in the
equilibrium position of the ship. Axes Ort and Oz are taken as the principal
axes. Vector - l l l is obtained by projecting principal axis Oz on plane O~(,
a perpendicular to this vector defines a unit base vector II of the nodal axis
which is an intersection of the principal planes O~( and Oxy. The angles
of rotation 'l/J and '!9 about axes Ort and ll, respectively, determine the trim
and the heel, whereas the angle of rotation cp about axis Oz , which is the
angle between II and Ox, determines the yaw. Expressions for the direction
cosines of the angles between the ship axes and the fixed ones are collected
2.4 Airplane angles and ship angles 59

in the following table

II II ~ 'Tl ( II
cos 'Ij; cos ip+ - cos ip sin 'Ij;+
sin ipcos{)
sin 'Ij; sin ip sin {) sin cp cos 'Ij; sin {)
- cos 'Ij; sin cp+ sin cp sin 'Ij;+
cos cp cos{)
sin 'Ij; cos cp sin {) cos cp cos 'Ij; sin {)
COSipcos{) - sin {) cos{)cos'lj;

Table 6 of the direction cosines for the ship angles

The ship and airplane axes possess the property that two angles (trim and
heel and correspondingly yaw and pitch) remain small for small changes
in the initial right angle between the principal axes. This choice has an
advantage over Euler's angles when only the nutation angle {) remains small
in the case of a small deviation of the moving principal axis from the
fixed one. If we take the airplane axes and view the axes O~ and Ox as
principal, then we can consider that only the angle between these axes
remain small for small deviations from the heading, whilst the angles of
yaw and pitch remain small for arbitrary roll values. Likewise, for a ship
we can view the axis Oz and the fixed vertical axis D( as the principal axes.
For Euler's angles only the nutation angle remains small whereas the angles
of precession and spin can be arbitrary. If we choose the ship axes, then
the angles of trim and heel are small and only the yaw angle is arbitrary.
Provided that the deviation of the body from its initial position is small,
all three angles describing airplane and ship angles remain small. If Euler's
angles are taken, then angle {) and the sum of angles 'Ij; + ip remain small for
a small deviation from the initial position. Indeed, as the angles between
the axes of the same name are small the diagonal elements in Table 2 differ
from unity only in second order terms, which implies the following

cos{)~l, cos('Ij;+cp)~l.

Table 2 takes the form

II II ~ 'Tl ( II
II x II 1 I 'Ij;+ip {) sin ip II
II y II - ('Ij; + cp) I 1 {) cos cp I"
II z II {) sin 'Ij; I -{)cos'lj; 1 II

Table 2* of the direction cosines


60 2. Rigid body kinematics - basic knowledge

Although angles 'IjJ and cP can take any values, the axes of the same name
remain close to each other. Tables 3 and 6 differ fundamentally from the
above table. The diagonal elements in these tables will differ from unity to
second order only if all three angles are small. In this case the non-diagonal
elements are of the first order of smallness. For instance, Table 3 takes the
form

II II ~ I 11 I ( II
II x II 1 I {} I -'IjJ II
II y II -{} I 1 I cP Ir
I z II 'IjJ I -cP I 1 II

Table 3* of the direction cosines

2.5 Using matrix multiplication to obtain tables of


direction cosines
Let us return to eq. (2.5) and study some simple cases. Given a rotation
through an angle CPI about axis ~ the table of direction cosines of the two
sets of axes Oxyz and 0~11( with respect to each other take the form

II II ~ I 11 ( II
II x 111 I 0 0 II
II y II 0 I cos CPI sinCPI Ir
II z II 0 I- sinCPI I cos CPI II
If we introduce the matrix corresponding to this rotation
100
al = 0 cos CPI sin CPI (2.5.1)
o - sin CPI cos CPI
the matrix form of the coordinate transformation for the case ~o = 0 is as
follows
(2.5.2)
The matrices for the rotation VJ2 about axis 011 and VJ3 about O( are
respectively given by
cos CP2 o - sin CP2 cos VJ3 sinCP3 0
o 1 o - sinCP3 cos VJ3 0 . (2.5.3)
sinCP2 0 COSCP2 o o 1
2.5 Using matrix multiplication to obtain tables of direction cosines 61

In the general case any rotation of a rigid body about a fixed point can
be achieved by three successive rotations about lines. For example, in the
case of Euler's angles, the first rotation is through an angle 'IjJ about axis
0(. This rotation brings the trihedron ~,7], ( into coincidence with axes
Xl, YI, Zl, with Zl, Xl and YI coinciding with axis (, the nodal axes and vec-
tor nl, respectively. The matrix form (2) for the coordinate transformation
is given by

Xl = fY-'Ij;~,

where fY-'Ij; is matrix fY-3 in which 'P3 is replaced by 'IjJ.


A further rotation {j about the nodal axis, that is axis OXI, puts the
trihedron OXIYIZI into orientation OX2Y2Z2, with axes OX2, OY2 and OZ2
coinciding respectively with OXI, n' and Oz. Since the rotation is performed
about axis OXI the rotation matrix fY-,J is obtained from matrix fY-I by
replacing 'PI with {j. The coordinate transformation in matrix form is

The third rotation makes the trihedron OX2Y2Z2 coincide with the body
axes Oxyz. The rotation matrix fY-<p is obtained from fY-3 under the assump-
tion that axis OZ3 remains fixed. Thus we have

Now we obtain

(2.5.4)
Thus, the rotation matrix fY- bringing axes 0~7]( into the final position
Oxyz is proved to be equal to the following matrix product

(2.5.5)
where fY-<p, fY-,J and fY-'Ij; are matrices of the type fY-3, fY-I and fY-3, respectively.
It is worthy of note that the matrices fY-<p, fY-,J and fY-'Ij; appear in the matrix
product in left-to-right sequence, i.e. the reverse to the rotation sequence.
Due to (1.13) it follows from (4) that

(2.5.6)
as the transpose of the product of two matrices is given by the product of
their transposes taken in the reverse order. It is clear that matrix (5) yields
Table 2 of direction cosines whilst matrix (6) yields the same table, with
x', Y' and z' being respectively replaced by ~,7] and (.
Table 3 for the direction cosines of the airplane angles is written as the
following matrix product

(2.5.7)
62 2. Rigid body kinematics - basic knowledge

where acp, au and a'IjJ are matrices of the type at, a3 and a2, respectively.
Table 6 for the direction cosines of the ship angles is given by

(2.5.8)
where acp, au and a'IjJ are matrices of the type a3, al and a2, respectively.
The rotation matrix which brings the fixed axes into coincidence with
the velocity axes is given by a matrix of the type (7)

(2.5.9)
The velocity axes Ox*y* z* are made coincide with the airplane axes by
means of the matrix

(2.5.10)
where a", and a{3 are matrices of the type a3 and al. Hence,

x = "Ix* = "Ia*~ = a~, "Ia* = a,

or

(2.5.11)
Notice that Table 4 offers "I'. The latter equation yields expressions for the
angles of heading, ascend and roll in terms of angles of yaw, pitch and roll,
as well as in terms of the angles of attack and slide. These formulae are
derived above under the assumption that angles a and f3 are small. In this
case "I' is given by Table 5 and the consequent equations in Sec. 2.4 can be
obtained by multiplication of the matrices.

2.6 Application to Cardan's suspension


2.6.1 Cardan's suspension
This facility, shown in Fig. 2.6, serves to fit a body to a fixed or moving
platform, e.g. ship or airplane, and consists of two gimbals, see [53] and
[42]. The bearings 1 and I' of the rotation axis of the outer gimbal (o.g.)
are fitted to the platform, whilst the inner gimbal (Lg.) is free to rotate
about the axis of the bearings 2 and 2'. The rotation axes of the outer and
inner gimbals are perpendicular to each other and intersect at the centre
o of both gimbals, this point remaining fixed relative to the platform. The
body (or a device) is fitted to the inner gimbals. In the case of a gyroscope
the inner gimbal carries the bearing of a spinning rotor. This is shown
schematically in Fig. 2.7a. The rotor axis is normal to the rotation axis
of the inner gimbal in the outer one and passes through the point 0 of
intersection of the gimbals' axes.
2.6 Application to Cardan's suspension 63

FIGURE 2.6.

The inner gimbal has two degrees of freedom relative to the platform. Its
position can be described by the rotation angle a of the outer gimbal and
angle fJ of rotation of the inner gimbal relative the outer one. Let O~TJ( be
the trihedron of the platform axes, O~ implying the rotation axis of the
outer gimbal. The axes Ox*y* z* of the inner gimbal coincide initially with
axes O~TJ(. Rotation a of the outer gimbal puts the rotation axis of the
inner gimbal and the normal to the outer gimbal plane into position Or/
(or Oy*) and 0(', respectively. In order to make this more illustrative Fig.
2.7b is rotated through 90° relative to Fig. 2.7a.
A further rotation through an angle fJ of the inner gimbal about axis
Oy* yields the trihedron position Ox*y* z*. Figure 2.7a displays the outer
gimbal rotated through an angle a, axis Oy* being directed normal to the
drawing away from the reader. The inner gimbal carries the rotor whose
bearings lie along the axis Oz*. Let Oxyz be the rotor trihedron (axes Oz
and Oz*coincide) which is obtained by rotating axes Ox* y* z* through an
angle cp about axis Oz*. The matrices of these rotations are as follows

1 0 0
ac> = 0 cos a sma
0 -sin a cos a

cosfJ 0 - sinfJ
a(3 = 0 1 0 (2.6.1)
sinfJ 0 cosfJ

cos cp sin cp 0
a<p = - sin cp coscp 0
0 0 1

Their elements are the direction cosines of the angles between the axes of
the rotated body and those of the platform. Matrix ac> describes rotation
of the outer gimbal relative to the platform, matrix a(3 describes rotation
64 2. Rigid body kinematics - basic knowledge

o.g.
a)
FIGURE 2.7.

of the inner gimbal in the outer one and , finally, matrix a", describes the
rotor rotation of the inner gimbal.
The matrix multiplication yields

cosf3 sin f3sin a - sinf3 coso.


o coso. sino. (2.6.2)
sin f3 - cos f3sin a cos f3cos a

This result is the table of direction cosines between two bases, namely O~'rJ(
and Ox*y* z*. What remains is to premultiply a{3aa by a", which results is
the following table

II II ~ 'rJ ( II
cos <p sin f3 sin 0'.+ - cos <p sin f3 cos 0'.+
cos <p cosf3
sin <pcosa sin<psina
- sin <p sin f3 sin 0'.+ sin <p sin f3 cos 0'.+
- sin <p cos f3
cos <p cos a cos <psina
sinf3 - cos f3 sin a cosf3 cos a

Table 7 of the direction cosines for the Candan's suspension

2.6.2 Double Cardan suspension


Two Cardan suspensions are mounted aboard a ship, their inner gimbals
being stabilised in the horizontal plane. The axis of the outer gimbals of the
2.6 Application to Cardan's suspension 65

!/

FIGURE 2.8.

first suspension coincides with the longitudinal axis Ox of the ship, whereas
the axis of the outer gimbals of the second suspension is directed along the
transverse axis Oy. The rotation angles of the outer gimbals relative to the
ship are designated as C¥1 and C¥2, see Fig. 2.8. The aim here is to obtain
the tables of direction cosines of the angles between the ship axes Oxyz
and the two sets of axes 01xiyizi and 02X2Y2Z2 of the inner gimbals in
terms of the above angles, [42].
The rotation matrices making axes Oxyz coincident with axes 01xiyi zi
and 02X2Y2z2 are C¥/31 C¥al and C¥/32C¥a2' respectively. The rotation angles /31
and /3 2 of the inner gimbals in the outer ones must be determined. The
first matrix is obtained from (6.2) by replacing c¥ and /3 with C¥1 and /31'
respectively. The second one is given by
1 0 o - sinc¥2
o cos/32 1 o (2.6.3)
o - sin/32 o
because the rotation through angle C¥2 is made about axis Oy. The matrix
multiplication yields
cos C¥2 0 - sin C¥2
C¥/32C¥a2 = sin/32 sinc¥2 cos/3 2 - sin/32 COSC¥2 (2.6.4)
cos /3 2 sin C¥2 - sin /3 2 cos /3 2 cos C¥2
Since the inner gimbals remains horizontal the third rows of this matrix
and Table 7, determining the direction cosines of the angles between the
vertical and the ship angles, are equal to each other, i.e.
sin/31 = cos/32 sinc¥2 , COS/31 sinc¥l = sin/32, cos/3 1 COSC¥l = cos/32 COSC¥2·
(2.6.5)
One of these equations is a sequence of the two others. Hence, we find
(2.6.6)
66 2. Rigid body kinematics - basic knowledge

or
. (3 - COS a l sina2 }
sm 1 - R '
. (3 - cos a2 sin al
(2.6.7)
sm 2 - R '
where

(2.6.8)
and E = ±1, the sign in all expressions (7) coincide due to the third equality
in eq. (5).
The rotation matrices a/31 aal and a/32aa2 take the form

cosa2 cos al sin al sin a2 cos 2 a2 cos a2


R R R
a/31 aal = 0 cos a l - sinal
cosal sina2 cos a2 sinal cosa2 cos a l
R R R
(2.6.9)
and
cosa2 0 - sina2
cos a2 sin al sin a2 cos a l cos 2 a2 sin al
af3 2 aa2 = R. R
R.
cosal sma2 cosa2 smal cos al cos a2
R R R
(2.6.10)
The third rows coincide whereas the others differ from each other. This
indicates that the axes of the inner gimbals are rotated relative to each
other about axis Olzi (02Z2'). Let"! denote the angle about which the
axes Olxiyi zi must be rotated unless they coincide with axes 02x2y2Z2'.
Then
cos,,! sin "! o
-sin,,! cos,,! o (2.6.11)
o o 1
By comparing the elements of the first rows we find
cosa2 cos,,!
O= R . . .
cosa2 = ~ cos,,!, cos a l smal sma2 + sm,,!cos a l,

. cos,,! 2. ..
sma2 = R cos al sma2 + sm"!smal·

Recalling expression (8) for R we obtain

cos,,! = R, sin,,! = - sinal sina2. (2.6.12)


2.6 Application to Cardan's suspension 67

Clearly, the elements of the second rows coincide.


The third column of Table 6 defines the direction cosines of the angles
between the upward vertical and the ship axes, however the same directions
cosines are given by the third row in matrix (9) or (10). This leads to the
following relations

tan 7jJ - sm
cos cp --{)- . cp tan {) = - tan a2, }
cos .f, (2.6.13)
. tan'f/
- sm cp--{)- - cos cp tan {) = tan al.
cos
When yaw is absent, then

(2.6.14)

which justifies the notation instrumental heel and instrumental trim for
-al and -a2, respectively. For small angles of yaw 7jJ ~ -a2. Notice that
tan7jJ = -tan,81 due to (6) and (14), that is the trim can be measured by
means of the rotation angle of the inner gimbal of the first suspension.

2.6.3 The platform on a Cardan suspension


A platform II is mounted on a moving base C, e.g. an airplane or a ship,
by means of Cardan's suspension. The bearings of the outer gimbal of the
suspension are fitted to one of the base axes (e.g. airplane or ship axes).
Those of the inner gimbal are fitted to the outer gimbal and finally the
platform bearings are mounted on the inner gimbal. These three axes are
mutually orthogonal and intersect at point O. The orientation of the two
sets of axes Oxyz and Oabc is assumed to be given in terms of the basis
(e.g. inertial or earth) axes O~T)(. Expressions for the direction cosines of
the angles between axes Oxyz and the basis axes in terms of the direction
cosines of the angles between axes Oabc and the basis axes, as well as the
rotation angles of the gimbals and the platform are sought.
Let II and C be the matrices describing the rotation of trihedron O~T)(
to Oabc and Oxyz, respectively. Then

a = II~, x = C~, a = IIC'x, (2.6.15)

that is matrix IIC' makes trihedron O~T)( coincide with Oabc. Let a denote
the rotation angle of the outer gimbal of the Cardan's suspension relative
to the base C. Let ,8 and I designate rotation angles of the inner gimbal in
the outer one and the platform in the inner gimbal, respectively. In passing
we note that the latter angle is denoted in Table 7 by cpo
Let OXY Z denote trihedron of the base. Its axes differ from the axes
Oxyz only in notation, for instance the fore-and-aft axis may be designated
as OZ. The matrix which makes axes OXYZ coincidental with Oxyz is
68 2. Rigid body kinematics - basic knowledge

z !I

o v c z

x z
FIGURE 2.9.

denoted by B. This matrix is called a matrix of notation alteration. Its


entries are zeros and ±l. For instance, this matrix is given by

010
B= 0 0 1
1 0 0

for the two sets of axes shown in Fig. 2.9.


Initially axis OX is the rotation axis for the suspension's outer gimbal,
whilst axes OY and OZ coincide with the rotation axes of the inner gimbal
and the platform. Thus the trihedrons OXY Z and Oabc initially coincide.
Provided that a denotes the matrix given by Table 7 of the direction
cosines we have

a = aX = a,a,6aaX, x = BX, a = a,a,6aaB' x, (2.6.16)

then, by virtue of (15) we arrive at the following relationship

(2.6.17)

Let us consider a platform stabilised in the system of base axes in such a


way that the axes of trihedron Oabc differ from axes O~Tf( only in notation.
Then II = B1 where B1 is a matrix of notation alteration. Then

(2.6.18)

Particularly, if a = j3 = 'Y = 0 then axes of the moving base coincide with


the base axes and

(C) a=,6=,=O = E = BB 1, C= BA'~ A'~A'~B'.


U~UfJU, (2.6.19)

Figures 2.10 and 2.11 show two variants of a drift indicator aboard a
ship, [53]. The first variant shown in Fig. 2.10 implies that B = E and
matrix C = a~a~a~ is the transpose of Table 7 of the direction cosines
2.6 Application to Cardan's suspension 69

u~x z

1L
o .z

FIGURE 2.10.

provided that 'P is replaced by ,,(, i.e.

cos "(cos (3 - sin "(cos (3 sin (3


cos "( sin (3 sin a - sin "( sin (3 sin a
- cos (3 sin a sin '13
c= +sin"(cosa +cos"(cosa
- cos "( sin (3 cos a sin "( sin (3 cos a
cos (3 cos a
+sin,,(sina + cos"( sin a
(2.6.20)

Comparing C with Table 6 yields the required relationships for the angles
of yaw, trim and heel in terms of ,,(, a and (3. As one can see they are rather
complicated. Figure 2.10 shows that 'P ::::; -,,(, '13 ::::; -a and 'l/J ::::; (3 only for
small angles.
For the second variant which is shown in Fig. 2.11 we have

o 0 -1 o 0 1
B = 0 1 0 B' = 0 1 0
1 0 0 -1 0 0

and, by virtue of (19), we obtain

- sin "( sin (3 cos a- cos"( sin (3 cos a-


cos (3 cos a
cos"( sina sin ,,(sina
C= - sin "( sin (3 sin a+ cos "( sin (3 sin a+
cos(3sina
cos"(cosa sin "(cos a
- sin (3 - sin "(cos (3 cos "(cos (3
(2.6.21 )
70 2. Rigid body kinematics - basic knowledge

tLy
J-.
:c

FIGURE 2.11.

For small angles we have <p = -a, 'IjJ = j3 and {) = 'Y . To explain the sign
of {) we notice that the reference axes for angles 'Y and {), i.e. Z and x, are
opposite in direction.
The relationships obtained simplify the consideration of more compli-
cated cases. For instance, let the platform be not absolutely stabilised and
its axes be obtained by rotation through small angles A, f..l and v about axes
03HZ. Then

1 v -f..l 1 v -f..l
~ = B 13 , a= -v 1 A ~, II= -v 1 A B'1
f..l -A 1 f..l -A 1

and relationship (17) yields

1 v -f..l
C = Ba'a a'(3 a''Y -v 1 A B~N. (2.6.22)
f..l -A 1

Assuming that for a perfectly stabilised platform a = j3 = 'Y = 0 and


axes of the moving base coincide with the base axes O~'T}( we obtain that
BB~ = E and B~ = B'. For small angles we arrive at the following equality

1 v-'Y
=B -v + 'Y 1 B' ,
f..l-f3 -A+a

in which matrix B is determined by the suspension design.


2.7 The velocity of a point in a rigid body 71

2.7 The velocity of a point in a rigid body


Differentiating vector equation (2.2) with respect to time and taking into
account that x', y' and z' are constants we arrive at the relationship

° °
r = ro + x ,di~
-
dt
+ y ,di~
dt
,di~
- +z -
dt '
(2.7.1)

where

(2.7.2)

represent the velocity vectors of points M and 0, respectively.


dO,
One can expand ;; in terms of the unit base vectors i~

(2.7.3)

where the elements wsr of the matrix W need to be determined. This matrix
is skew-symmetric. Indeed, differentiating the identity

i~ . i~ = 8sr,

yields

di~
-'1
0' di~
=--'1
0'
dt r dt s·

If follows from (3) that the right-hand side of the latter equation is equal
to -wsr whereas the left-hand side is w rs , which completes the proof.
Representing the skew-symmetric matrix W in the form

o
W= (2.7.4)

cf. (A.2.3), we can rewrite formulae (3) in a different form

(2.7.5)

Here w denotes the vector which is referred to as the angular velocity vec-
tor, WI, W2 and W3 being the projections of vector w on the axes of the
72 2. Rigid body kinematics - basic knowledge

coordinate system Ox' y' z'. In Sec. 2.12 we will show that these three val-
ues are transformed under a change of the coordinate basis as components
of the vector. Equation (5) can be recast in the form

(2.7.6)

thus

d" 3
" •
It ""di
Is = '""' .
~ cttmctsm = EstqWq ,
m=l

which yields expressions for the angular velocity projections in terms of the
direction cosines and their time derivatives

(2.7.7)

For example, the full expression for WI is

and so on.
By virtue of eq. (5), formula (1) for the velocity distribution in a rigid
body takes the form

v = Vo + wx (x'i~ + y'i~ + z'i~) ,


or

v = Vo +w x r' = Vo + wx (r - ro), (2.7.8)

where the second term expresses time derivative of vector r' = r - ro

dr'"
""di = r = w x r , = w x (r - ro ) . (2.7.9)

2.8 Vector of infinitesimal rotation


We consider a rigid body having a fixed point O. The body is assumed to
move from a given position to another which is infinitesimally close to the
given position. The vector of an infinitesimal displacement Or of a point M
denoted by position vector r = CJJI1 is normal to r. This follows from the
equation
1 r2
r . or =8-r . r =0 - = 0
2 2
2.8 Vector of infinitesimal rotation 73

()

FIGURE 2.12.

because the length of vector r is invariable in the rigid body. Thus one can
represent vector 8r as follows

8r = (J x r, (2.8.1)

where (J is an infinitesimal vector identical for all points of the body. Indeed,
if one considered this statement to be incorrect and wrote for two points
in the body the following equations

one would have

The left-hand side is a variation of the scalar product rl . r2. This variation
is equal to zero since the length of vectors rl and r2 as well as the angle
between them do not change. For this reason

implying that (Jl - (J2 = 0 since vector rl x r2 is arbitrary.


Equation (1) suggests that

8r = I(JI r sin Ct.

With the help of Fig. 2.12 we see that the absolute value 181 of vector
(J expresses the angle of infinitesimal rotation for any position vector r.
Therefore (J is called the vector of infinitesimal rotation. The definition of
the vector of finite rotation will be given in Sec. 3.l.
Provided that a rigid body has a fixed point 0 we have Va = 0 and r'= r.
Equation (7.8) for the absolute velocity takes the form

V = w x r. (2.8.2)
74 2. Rigid body kinematics - basic knowledge

With this in view and setting 8r = vdt we obtain, by comparing eqs. (1)
and (2),

() = wdt. (2.8.3)
This equation relates the above introduced vector w with the vector of
infinitesimal rotation () which explains why w is referred to as the vector
of angular velocity.
Let the rigid body with an immovable point be subject to two successive
infinitesimal rotations (}l and (}2. After the first rotation the position vector
r becomes

r' = r + (}l x r,
whilst after the second rotation it becomes
r"= r'+(}2 x r' = r + (}l x r + (}2 x (r + (}l x r).
It would not only be redundant but wrong to keep the small values of the
second order (}2 x ((}l x r) because we neglected the values of this order in
the original definitions. Thus we obtain

r" - r = ((}l + (}2) x r = () x r,

where () denotes an infinitesimal rotation corresponding to the total dis-


placement 8r = r" - r. Therefore

(2.8.4)

and the infinitesimal rotations are commutative. By virtue of relationship


(3) this statement is also correct for angular velocities.

2.9 Angular velocity vector in terms of the time


derivative of Euler's angles
Let us consider two positions of the body with a point fixed in space. The
first position is defined by Euler's angles 1/;, {} and cp whereas the second
position, being infinitesimally close to the first one, is given by 1/;+81/;, {}+8{}
and cp + 8cp. The body can be moved from the first position to the second
one by means of three infinitesimal rotations whose vectors are given by
the following expressions

(}l = h81/;, (}2 = n8{), (}3 = i~8cp.


In accordance with eq. (8.4) the resultant infinitesimal rotation is just a
geometrical sum of the components

() = h81/; + n8{) + i~8cp. (2.9.1)


2.9 Angular velocity vector in terms of the time derivative of Euler's angles 75

By meanS of (8.3) we obtain the following expression for the angular veloc-
ity vector
w = i3~ + nlJ + i~cp. (2.9.2)
Its projections on the axes that are fixed in the body are

WI = W . i~ = ~ sin 1'J sin rp + ~ cos rp, }


W2 = W . i; = 1j; sin 1'J cos rp - 1'J sin rp, (2.9.3)
W3 = W . i~ = ~ cos 1'J + cp.
Projections of W on the axes fixed in space are designated as WI, W2 and
W3, and are given by
WI = W . il = lJ cos 1j; + cp sin 1'J sin 1j;, }
W2 = W . i2 = lJ sin 1j; - cp sin 1'J cos 1j;, (2.9.4)
W3 = W· i3 = ~ + cpcos1'J.
Analogous equations can be constructed for projections of the infinitesi-
mal rotation vector. To this end, it is sufficient to replace the time deriva-
tives~, lJ and cp in eqs. (3) and (4) by variations 81j;, 81'J and 8rp, respectively.
Then One obtains the following expressions for the projections of () on the
axes that are fixed in the body
(h = sin 1'J sin rp81j; + cos rp81'J, }
e2 = sin 1'J cos rp81j; - sin rp81'J, (2.9.5)
e3 = cos 1'J81j; + 8rp,
whilst the projections on the axes that are fixed in space are

~l = sin 1'J sin 1j;8rp + cos 1j;81'J, }


e2 = - sin 1'J cos 1j;8rp + sin 1j;81'J, (2.9.6)
fh = cos 1'J8rp + 81j;.
We notice at this point that the angular velocities of the "half-fixed" tri-
hedron n, n l , h and the "half-moving" trihedron n, ni, i~ are equal to ~h
and ~h + nlJ, respectively. Due to formulae (7.5) we find

n = ~h x n = ~nl' nl = -1j;n, ~: =0 (2.9.7)

and
n = ~nl = ~ (n' cos 1'J - i~ sin 1'J), }
ll' = -~n cos 1'J + lJi~, (2.9.8)
di'· .
~ = 1j;n sin 1'J - 1'Jn'.
dt
Expressions for the variations 8n,8nl, 8n' and 8i~, which are infinitesimal
displacements of ends of the corresponding vectors n, n l , n' and i~ subject
to infinitesimal rotation (), are obtained by replacing ~, lJ and cp in eqs. (7)
and (8) with variations 81j;, 81'J and 8rp.
76 2. Rigid body kinematics - basic knowledge

2.10 Calculation of three-index symbols


As indicated in Sec. 1.5 when considering a rigid body motion it is natural
to view projections Ws of the angular velocity vector and projections VOs
of the velocity vector on the axes fixed in the body as quasi-velocities. We
keep the notation WI, w2 and W3 for the first group of quasi-velocities and
denote the projections VOs by W4, W5 and W6. This ensures that there are
~ .6 2 . 5 = 90 three-index symbols to be calculated.
1. Quasi-velocities of the first group are related to the generalised veloc-
ities by means of formulae (9.3). The inverse relationships are as follows

'Ij;. =
. SIn 'U 1 (WI sin <p + W2 cos <p) ,
--:--::a }
1) = WI cos <p - W2 sin <p, (2.10.1)
if = - (WI sin <p + W2 cos <p) cot 1) + W3.

Variations of the quasi-coordinates 87r1, 87r2 and 87r3 are equal to pro-
jections of the infinitesimal rotation vector on the axes fixed in the body.
Their expressions in terms of variations 8'1j;,81) and 8'1j; of the generalised
coordinates are defined by relationships (9.5). We obtain the inverse ex-
pressions from eq. (1) by replacing ;P,fJ,if and Ws with 8'1j;, 81), 8'1j; and Bs,
respectively.
We start with the following equation

Accounting for the "rule d8 = 8d" for calculating the difference between
the above quantities we obtain

( fJ8'1j; - ;P81) ) cos 1) sin <p +

(if8'1j; - ;P8<p) sin 1) cos <p + (fJ8<p - if81) ) sin <po

In this equation it is essential that the generalised velocities and varia-


tions of the generalised coordinates be replaced by their values due to eq.
(1) in terms of the quasi-velocities and variations of the quasi-coordinates,
respectively. On performing this for the differences

we arrive at the formulae

(87rIt - 8WI = w387r2 - w2 87r 3, }


(87r2t - 8W2 = W187r3 - W3 87r 1, (2.10.2)
(87r3t - 8W3 = W287r1 - WI 87r2·
2.10 Calculation of three-index symbols 77

By virtue of (8.3) we have


1
'32 = -'23 =
1 1
, (2.10.3)
The remaining symbols with indices 1,2 and 3 are identically equal to zero.
2. One could perform a similar calculation based upon the formulae

etc. which are obtained by projecting eq. (7.2) for the velocity vector of
pole 0 on the axes fixed in the body. This calculation would be very cum-
bersome. It can be simplified by using the relationships

(2.10.4)
as well as the formulae for differentiating (7.5) the unit vectors and the
following equations for variations

8i~ = 0 x i~. (2.10.5)


In eq. (4) 8ro stands for the vector of infinitesimal displacement of the
pole. Let us recall that projections of 0 on the axes fixed in the body are
denoted by 87r s. Then we have

(87r3+st = (brot . i~ + bro· (w x i~), bW3+s = bvo . i~ + Vo . (0 x i~).


Taking into account relationship (1.7.10) we find

(b7r3+s)- - 8w3+s = (bro x w - Vo x 0) . i~. (2.10.6)


By setting s = 1,2,3 we obtain the equations

(b7r4t - bW4 = W3b7r5 - W287r6 - (W587r3 - W6b7r2), }


(b7r5t - bW5 = Wlb7r6 - W3b7r4 - (W6b7rl - W4 b7r3), (2.10.7)
(b7r6t - bW6 = W2b7r4 - Wl b7r5 - (W4 b7r2 - W5 b7r d·

Let us recall that W4,W5,W6 and b7r4,b7r5,b7r6 denote the projections of


vectors Vo and bro on the axes fixed in the body, respectively. The following
twelve values are non-trivial
,j5 = -,g3 = 1, '~2 = -'~6 = 1, }
,f6 = -'~l = 1, '~3 = -'~4 = 1, (2.10.8)
'~4 = -'~2 = 1, '~l = -'~5 = 1.
3. Provided that the projections Ws of the angular velocity are taken as
quasi-velocities and using eq. (9.4) and the inverse equations in the form

(2.10.9)
78 2. Rigid body kinematics - basic knowledge

we arrive, by repeating the above calculation, at the following relationships


(87rlt - 8Wl = w287r3 - W3 87r 2, }
(87r2t - 8w2 = W387rl - WI 87r3, (2.10.10)
(87r3)- - 8w3 = W187r2 - W287rl.
The three-index symbols are as follows
1
'Y23 1 1
= -'Y32 = ,
2
'Y31 = -'Y13
2
= 1, (2.10.11)
They differ from the above symbols (3) corresponding to the case in which
the quasi-velocities are projections on the axes fixed in the body only in
sign.
4. Using a trihedron of moving axes which are not fixed in the body can
often simplify the construction of equations of motion. The "half-moving"
trihedron n, n', i; introduced in Sec. 2.3 can serve as an example. The
angular velocity vector of the "half-moving" trihedron is given by
w' = {}n + ;Pi3 = -an + ;p sin {)n' + ;p cos {)i~ . (2.10.12)
The angular velocity vector of the body is given by
w = w' + i~<p. (2.10.13)
Let us take projections of won the axes of the "half-moving" trihedron as
quasi-velocities
WI = -a, W2 = ;p sin {), W3 = ;p cos {) + <p. (2.10.14)
The variations of the quasi-coordinates are then
87fl = 8{), 87f2 = sin {)8,¢, 87f3 = 8cp + 8'¢ cos {). (2.10.15)
Only the following three-index symbols do not vanish
= 1.
2 2 {) 3 3 (2.10.16)
'Y12 = -'Y21 = cot , 'Y21 = -'Y12

The three-index symbols with superscript 1 are equal to zero since {) is a


"true" coordinate.
5. Let us now take projections of the velocity vector Vo of a pole 0 on
axes of the" half-moving" trihedron n, nl' h
w4=vo·n=xocos'¢+Yosin,¢, }
W5= ~o . nl = -xo sin'¢ + Yo cos ,¢, (2.10.17)
W6 = Zo°

The latter expression is integrable and the three-index symbols with super-
script 6 are equal to zero. As the quasi-velocity W2 is given by eq. (14) we
have
- (;P8x o - x o8'¢ ) sin'¢ + (;P8yo - yo8'¢ ) cos '¢
. 1
'¢87f5 - w58'¢ = --;--{) (w287f5 - w5 87f2)
sm
2.10 Calculation of three-index symbols 79

and we obtain
4 4 1 5 5 1
125 = -'52 = sin '19' 142 = -'24 = - . - . (2.10.18)
sm'19
A sphere and a ring rolling without slippage on a fixed plane are classical
examples of systems subject to non-holonomic constraints.

2.10.1 Sphere rolling on a rough plane


The position of the sphere is given by five parameters: coordinates xo, Yo of
the centre of the sphere in the fixed coordinate system Oxyz and three Euler
angles. Axis Oz described by the unit vector h is normal to the plane on
which the sphere rolls. The equations of non-holomonic constraints express
the condition of vanishing velocity at the point of contact P. Using eq. (7.8)
we obtain

as rp - ro = -ha, a being the sphere radius. As above, Ws are the projec-


tions of the angular velocity vector on the fixed axes.
Equations for the non-holonomic constraints are

Xo - aW2 = 0, Yo + aWl = O.
The system has three degrees of freedom. Let us take WI, W2, W3 as quasi-
velocities and introduce

W4 = Xo - aW2 = Xo - a (~sin 'ljI - cp sin '19 cos 'ljI), }


(2.10.19)
W5 = Yo + aWl = Yo + a ( ~ cos 'ljI + cp sill'!9 sin 'lj! ) .

Due to equations for the non-holonomic constraints, W4 = 0, W5 = O. Then


we have

and, by virtue of eq. (10), we obtain

It3 = -'~l = a, 1~3 = -'~2 = a. (2.10.20)

The remaining three-index symbols are determined by eq. (10). The gen-
eral problem of a body with an arbitrary convex surface rolling on a fixed
plane is considered in Sec. 2.12.

2.10.2 A ring rolling on a plane


The position of the ring is described by the same five parameters Xo, Yo, 'ljI, '19
and t.p as in the case of the sphere. They are shown in Fig. 2.13. The
80 2. Rigid body kinematics - basic knowledge

-;r----n,

p p

FIGURE 2.13.

coordinate Zo of the centre of the ring is expressed in terms of angle {} by


means of the following relationship

Z = a sin {}. (2.10.21)

In the case being considered

rp - ro = an', (2.10.22)

where a is the ring radius and n' denotes the unit vector of the half-moving
trihedron n, n' , i~ directed upward along the steepest line of the ring plane.
The velocity of the contact point P is given by

Vp Vo +w x (rp - ro) = Vo - aw x n'


Vo - a (nwl + n'w2 + i~W3) x n'.

While deriving the latter result we used eq. (12) and adopted notation (14)
for the quasi-velocities. When slipping is prevented we have

Vp = Vo - ai~wl + anW3 = O. (2.10.23)


We write equations for the non-holonomic constraints by projecting this
relationship on axes of the" half-fixed" system n, n'. Making use of notation
(17) we obtain
W4 + aw3 = 0, W5 + aWl sin{} = O. (2.10.24)

Projecting on axis Oz yields the following integrable relationship

W6-aWlcos{}=0 or zo=a7Jcos{}, (2.10.25)

which is a sequel of eq. (21). Instead of using W4 and W5 it is convenient to


introduce new quasi-velocities

(2.10.26)
2.11 Acceleration of a point in a rigid body 81

which vanish due to equations for the non-holonomic constraints. Therefore,


five quasi-velocities being used are the three quantities given by eq. (14)
and the two quantities introduced by eq. (26).
The three-index symbols corresponding to the first group are given by
eq. (16). Further we have, due to eqs. (18) and (2)

(D1f~t - DW~ = ~
Slnu
(w2D1f5 - w5D1f2) + a (W2D1fl - WID1f2).

Replacing here D1f5 and W5 by their values from (26) we obtain

(D1f~t - Dw~ = ~
Slnu
(W2D1f; - W;D1f2) . (2.10.27)

Noticing that expression WI sin {) = ~ sin {) is integrable we obtain, due to


eq. (18),

or replacing D1f4 and W4 by their values from (18) we have

(D1f;)- - DW;; = .1 .0
Slnu
(w2D1f~ - W~D1f2) + smu
.a (w2D1f3 -
.0 w3D1f2) .

Hence, the three-index symbols are as follows

4 4 1
'Y25 = -'Y52 = ~, (2.10.28)
smu

5 5 1 5 5 a (2.10.29)
'Y42 = -'Y24 =~,
Slnu 'Y23 = -'Y32 = sin {).

2.11 Acceleration of a point in a rigid body


Differentiating expression (7.8) for velocity of a point in the rigid body with
respect to time yields the acceleration vector w = v. Taking into account
eq. (7.9) we arrive at the following equation

w = Wo + w x r' +w x (w x r'). (2.11.1)

Here Wo = vo designates accelerations of the pole and r' = r - ro is the


position vector with respect to the pole. The time derivative of the angular
velocity vector w is referred as to the vector of angular acceleration and is
denoted by
3
e =W= L (W 8 i~) - .
8=1
82 2. Rigid body kinematics - basic knowledge

Applying the formulae for differentiating unit base vectors (7.5) we obtain
3 3 3 3
e = LW8i~ + LWsW X i~ = LWsi~ +W X w = LW8i~. (2.11.2)
8=1 8=1 8=1 8=1

Thus projections C8 of the angular acceleration vector of the rigid body


on the axes fixed in the body are equal to the time derivatives of the
projections of the angular velocity vector on these axes. Expressions for Cs
in terms of Euler's angles and their first and second time derivatives are
obtained by differentiating equalities (9.3) with respect to time. Clearly,
projections f;s of vector e on the axes fixed in space are also equal to the
time derivative of projections Ws of vector w on these axes.
We consider now a trihedron of moving axes which are not fixed in the
body and denote the unit vectors ofthis trihedron by i;. Let w' and = W· w;
i; designate the vector of angular velocity of this trihedron and projections
of the angular velocity vector on axes i;, respectively. Then, repeating the
above calculation we obtain
3 3
e = W = """'
~S8 w*i* + w' X w = """'
L-tS8w*i* - (w - w') X W . (2.11.3)
s=1 8=1

Here w - w' is interpreted as the angular velocity of the body with respect
to trihedron i;. For example, in the case of the "half-moving" trihedron
n, n ' and i~ we have

e = On + (~sin '!9 r n' + (~cos '!9 + <P r i~ -

[ll'l9 + n' ~ sin '!9 + i~ ( ~ cos '!9 + <p)] = (0 + <p~ sin '!9) n +
i~ <P X
(~sin '!9 + ~~cos'!9 - <p~) n' + (~cos'!9 - ~~sin '!9 + cp) i~. (2.11.4)
In formula (2) for the acceleration in the rigid body the second term
e r' is referred to as the rotational acceleration whilst the third term
X
w X (w x r') is named the centripetal acceleration. The absolute value for
the centripetal acceleration is
(2.11.5)
where h is the distance between the rotation axis passing through 0 and the
point under consideration. This acceleration is directed toward the rotation
axis.
Expression for the centripetal acceleration can also be cast in the follow-
ing form
w x (w x r') = w (w· r') - w 2 r' = (ww - Ew· w) . r'. (2.11.6)
Here ww is a dyadic and E is the unity tensor. The basics of the dyadic
calculus is briefly outlined in Sec. 4.3.
2.12 Matrix form for velocity and acceleration in a rigid body 83

2.12 Matrix form for velocity and acceleration in a


rigid body
To begin with, let us recall the notation introduced in Sec. 2.2. Row-matrix
of the coordinates of a point in the coordinate axes Of,"7( fixed in space is
denoted by f" that in the coordinate axes Oxyz fixed in the body is denoted
by x, and f,o denotes the row-matrix of the coordinate of the pole O. The
rotation matrix which aligns the trihedron Of,"7( with trihedron Oxyz is
designated as a. Its elements are the direction cosines of the angles between
the axes of two coordinate systems, namely ask is the cosine of the angle
between the s - th axis of trihedron Oxyz and the k - th axis of trihedron
Of,"7(.
Formulae for transformation of coordinates which are components of the
position vector are written in the form
(2.12.1)
Similarly, one can construct formulae for the transformation of components
of any vector a. Let a and a denote column-matrices whose elements are
the projections as on the fixed axes and as on the axes Oxyz, respectively,
i.e.
a- ,
= a a a =
-
aa. (2.12.2)
Taking the derivative of (1) with respect to time yields
~-~o=iix, (2.12.3)

since x does not depend on time. Further, ~ - ~o represents column-matrix


whose entries v - Vo are projections of vector v - Vo on axes Of,"7(. Due to
eq. (2) we obtain

v - Vo = a (v - vo) = a (~ - ~o) = aa/ x. (2.12.4)

This formula gives velocity of points in the rigid body. Matrix aa' is skew-
symmetric which follows from the equation
., = -aa,
(aa. ')' = aa .,
in which the latter equality is proved by differentiating relationship aa' =
E with respect to time.
Therefore one can adopt
w= aa', (2.12.5)
where the skew-symmetric matrix
W11 = 0 W12 = -W3 W13 = W2
w= W21 = W3 W22 = 0 W23 = -WI (2.12.6)
W31 = -W2 W32 = WI W33 = 0
84 2. Rigid body kinematics - basic knowledge

is given by the three following values

Using the Levi-Civita symbols the latter can be cast in a short form

Wr = -ErstWst (r = 1,2,3) . (2.12.7)

It will be shown below that Wr can be treated as components of the


angular velocity vector w. Equation (4) takes the form

v - Va = WX (2.12.8)

and presents a matrix form of the vector equation for velocity of a point
in the rigid body (7.8) provided that this equation is projected on the axes
fixed in the body. Replacing in eq. (8) column x by expression (1) and
premultiplying both sides by a', we arrive by virtue of (2) to the following
relationship

(2.12.9)

which is the result of projecting the vector equation for the velocity of a
point in the rigid body on the fixed axes. The skew-symmetric matrix

w= a'wa = a'aa'a = a'a (2.12.10)

determines the column-matrix wr of projections of vector W on axes fixed


in space. Analogous to (7) we have

(2.12.11)

It is necessary to prove that two trihedrons wr and Ws transform like


components of a vector under a coordinate transformation, i.e.
3
Ws = 2:asrwr.
r=l

Due to (10) we have

W= awa'. (2.12.12)

Multiplying the matrices we obtain

Referring to eq. (1.9) this can be rewritten in the form


2.12 Matrix form for velocity and acceleration in a rigid body 85

or
3
Wst = -EstrWr = -Estr L O:rqWq ,
q=l

thus the statement is proved.


Defining the position of the rigid body by Euler's angles and presenting
the rotation matrix 0: in the form of eq. (5.5), we can write the angular
velocity matrix (5) as follows

Here, in accordance with the definition of matrices 0:'1" o:,} and 0:1j;

- sin'P - cos 'P 0 o -1 0


./
0:
'I'
= cos'p - sin'P 0 100
o o 0 000

etc. Applying eq. (13) we obtain

0 -1 0 0 0 0
w=rp 1 0 0 + iJo:'I' 0 0 -1 /
0:'1'
0 0 0 0 1 0
0 -1 0
+ 'ljJ0:'I'0:,} 1 0 0 /
O:,}O:'I'
/

0 0 0

o ;p sin {) cos 'P


-iJ sin 'P

o -;p sin {) sin 'P


-iJcos 'P
-;p sin {) cos 'P ;p sin {) sin 'P o
+iJsin 'P +iJ cos 'P
(2.12.14)

which, due to eq. (7), is equivalent to formulae (9.3). Matrix is calculated w


by analogy.
We proceed now to formulae for the acceleration of a point in the rigid
body. By virtue of eqs. (9) and (10) we have

(2.12.15)

Differentiating this equality with respect to time yields

(2.12.16)
86 2. Rigid body kinematics - basic knowledge

where

(2.12.17)

is a skew-symmetric matrix constructed by rule (11) from projections c: s of


the angular acceleration vector on the fixed axes. We also have

0 -W3 W2
2
-w-2 + WI
-2
WlW2 WlW3
W3 0 -WI W2Wl -w-2 +W 2 W2 W3
-2

-W2 WI 0 W3Wl W3W2 -W-2 + W3


-2

-Ew'w+ww',
where W' is the row-matrix WI, W2, W3 I II. This relationship follows
immediately from eq. (A.4.7). Therefore,

(2.12.18)

This is a matrix form of the vector formula for acceleration in the rigid
body (11.1) provided that this formula is projected on the fixed axes.
Let w - Wo denote the column-matrix of projections of the vector w - Wo
on the axes fixed in the body. To regard w - Wo as v - Vo would be a grave
error. By virtue of (2) and (1) we have

w-wo=a(~-~o) =a(ww'-Ew'w+~)a'x. (2.12.19)

But aw = w, W' a' = w' etc. and hence due to eqs. (10) and (5)

a~a' a (f;;) - a' = a (a'wa)- a' = aa'w + (w)- + waa'


w2 + (w r + w(w)' = (w r = t,
since w = - (w)' is a skew-symmetric matrix. Hence

(w-)- = c:- = ;::;, =


ac:a -
a (;::;)
w a. , (2.12.20)

This expression, being analogous to eq. (17), expresses the skew-symmetric


matrix of angular acceleration t in terms of projections of the angular
acceleration vector e on axes fixed in the body. Along with the above result
it is shown that projections of the angular acceleration vector on these axes
are equal to the time derivatives of the corresponding projections of the
angular velocity vector w.
Equation (19) takes the form

w - Wo = (ww' - Ew' w + t) x = (w 2 + t) x (2.12.21)

and represents a matrix form of the vector formulae for acceleration of a


point in the rigid body in projections on axes fixed in the body.
2.13 Differentiation of vector in a moving coordinate system 87

To closing this Section we express the quantities in eq. (21) in terms of


the rotation matrix, such that
(2.12.22)
and
(w_)2 =
_, _ .,. ,
-w w = -aa aa = -aa .
.. ,
(2.12.23)
Thus we obtain
.. ,
W - Wo = aa x, (2.12.24)
where w - Wo denotes the column-matrix of projections of vector w - Wo
on the axes fixed in the body. In view of (2) we find
c
<, - C=' "
<'0 ax. (2.12.25)
This can also be obtained by directly differentiating eq. (1).

2.13 Differentiation of vector in a moving


coordinate system
We consider a vector a whose projections on the trihedron of unit vectors
i~, i~, i~, rotating with angular velocity w, are al,a2,a3, respectively. In this
case
(2.13.1)
While differentiating this vector we should account for formulae (7.5) for
derivatives of the unit vectors. Thus we obtain
it ali~ + a2i~ + a3i~ + alw x i~ + a2W x i~ + a3w x i~
ali~ + a2i~ + a3i~ + w x (ali~ + a2i~ + a3i~).

Let;; denote a vector whose projections on the trihedron i~, i~, i~ are as,
that is
(2.13.2)
Equation (2) then takes the form

it =;; +w x a. (2.13.3)
Vector it can be referred to as the relative or local derivative of a, the term
w x a takes into account rotation of the trihedron i~, i~, i~ on which the
vector is projected.
Two special cases of formula (3) were described above. In eq. (7.9) pro-
jections of r' on axes Ox' y' z' are constant and thus the first term in eq.
(3) vanishes. In eq. (11.2) vector w is differentiated and the second term in
eq. (3) vanishes.
88 2. Rigid body kinematics - basic knowledge

2.14 Relative motion


Let us consider motion of a system of particles S with respect to moving
axes Ox' y' z'. The constraint equations in this coordinate system do not
contain time explicitly

(k=l, ... ,r). (2.14.1)

The coordinates x~, y~, z~ of any point M of the system S can be expressed
in terms of n = 3N - r independent generalised coordinates by means of
the following relationships

(2.14.2)

or, in vector form,

(2.14.3)

Notice that these relationships do not contain time explicitly. In what fol-
lows, subscript i is omitted.
The position of axes Ox' y' z' is prescribed by coordinates Xo, Yo, Zo of its
origin (referred to as the pole in what follows) in the system of inertial
axes Oxyz and three angular coordinates, for example Euler's angles. The
direction cosines of Table 1 in Sec. 2.1 are expressed in terms of Euler's
angles. We can determine the position vector r of the material point with
respect to the origin of the inertial coordinate system by means of eq.
(2.2) whilst eq. (2.4) gives expressions for x', y' and z' in terms of x, y, z
and six parameters describing the motion of trihedron Ox' y' z'. Substituting
these equations into (1) we arrive at the constraint equations in the inertial
coordinate system. It is necessary to distinguish between two cases. In
the first case, rotation of trihedron Ox' y' z' is prescribed and then time t
appears in the constraint equations by means of the relationships

Xo = Xo (t), Yo = Yo (t) , Zo = Zo (t), }


(2.14.4)
'Ij;='Ij;(t), {)={)(t), <p=<p(t),

determining the pole and angular coordinates. In this case the constraints
are non-stationary and the number of degrees of freedom is n. In the second
case, motion of trihedron Ox' y' z' is not prescribed, then 3N coordinates
of the point in the inertial coordinate system and six parameters defining
the position of trihedron Ox'y'z' appear in the above r equations. In this
case the constraints are stationary and the number of degrees of freedom
is n + 6. It is clear that there may also be intermediate cases where motion
2.14 Relative motion 89

of trihedron Ox' y' z' is only partially prescribed, for example, only pole
motion is prescribed with the angular coordinates not being prescribed.
Let us imagine that a certain rigid body is bound with the coordinate
system Ox' y' z' and consider point M* of this body coinciding with point
M of system S instantaneously. It is clear that the velocity and the accel-
eration of point M* are determined through rigid body kinematics, i.e. by
eqs. (7.8) and (11.1). They are referred to as the translational velocity and
translational acceleration and denoted by Ve and We, respectively. If it is
necessary the angular velocity vector wand the angular acceleration vector
e of the body bound with trihedron Ox'y' z' will be denoted by a subscript
e.
Imagine now that the coordinate system Ox' y' z' does not move. In other
words, being constantly bound with these axes we observe the movement
of the points of system M. The velocity and acceleration of point M deter-
mined under this condition are called the relative velocity and the relative
acceleration and are denoted by Vr and Wr, respectively. As pointed out in
Sec. 1.3 they are calculated by differentiating relationship (2), or its vector
counterpart (3), with respect to time

(2.14.5)

An observer in the system of inertial axes Oxyz determines the absolute


velocity Va and the absolute acceleration Wa of point M in system S as
follows

Va = r, Wa =r= Va. (2.14.6)

Alternatively, differentiating eq. (2.2)

r = ro + r' = ro + x'i~ + y'i; + z'i~, (2.14.7)

with respect to time and repeating the derivation of formulae (7.8) under
the condition that the position vector r' is not invariable with respect to
axes Ox' y' z', we arrive at the following result

Va = Vo +w X r' + i~ X' + i; y' + i~ Z' = Vo +w X r' + r*' , (2.14.8)

which can be also obtained from eq. (13.3) by replacing a with r'.
Noticing that

(2.14.9)

we arrive at the theorem of addition of velocities

Va = Vo +w X r' + Vr = Ve + Vr , (2.14.10)
90 2. Rigid body kinematics - basic knowledge

which expresses the absolute velocity as the geometric sum of the transla-
tional and relative velocities.
Since vectors r~ and i~ do not depend upon the generalised coordinates
and r' is independent of the generalised velocities, it follows from eqs. (4),
(8) and (10) that

OVa
(s = 1, ... ,n). (2.14.11)
oqs
Differentiating eq. (10) with respect to time leads to the theorem of
addition of accelerations. In order to prove the theorem we should repeat
the derivation of formula for acceleration in the rigid body (13.3) for vectors
r' and v r . We obtain

Wa = Wo + ex r' + wx (w x r') + wx r*' + Vr


* +w x v r .

In view of eqs. (9) and (5)

r*' = v r , v* r = W r. (2.14.12)

Therefore

Wa = Wo + ex r' + wx (w x r') + Wr + 2w x v r , (2.14.13)

or alternatively

Wa = We +Wr +WCor· (2.14.14)

Absolute acceleration is equal to the geometric sum of the translational


acceleration We, the relative acceleration W r and the Coriolis acceleration
WCor. The latter is given by the formula

WCor = 2w x v r • (2.14.15)

We proceed now to the formula which is important for the forthcoming.


Let a denote a vector depending on generalised coordinates qt, ... , qn, then

• *
a = a +w x a = L...J
~oa.
a qs + w x a.
s=1 qs
(2.14.16)

In particular,

(2.14.17)
2.15 Absolute acceleration of point moving over the rotating earth 91

2.15 Absolute acceleration of point moving over


the rotating earth
Let v N, Va and Vh denote components of the velocity vector of a mate-
rial point (e.g. an airplane) relative to the rotating earth in the northern,
eastern and vertical directions, respectively. The earth is assumed to be a
perfect sphere. Equations taking into account that the earth is a spheroid
are given in [14]. We introduce a geocentric system of axes Oxyz. The axis
z is directed upwards and coincides with the local vertical. Axis x is tan-
gent to the meridian pointing north whilst axis y is tangent to the parallel
pointing west. This system of axes rotates with the earth and moves on the
surface of the earth in such a way that the point in question remains on
axes Oz at any time instant. The latitude and the longitude of the point
are designated as <I> and ,x, respectively, and h denotes the height above
the origin 0 of the axes system Oxyz.
Denoting the radius of the earth as R we have

(2.15.1)

The components of the velocity vector v of the geocentric system Oxyz


along the axes of this system are equal to

(2.15.2)

whereas the components of the angular velocity are given by

WI = ,x +
U cos
( .) +h
<I> = RVa + U cos <I> , )
. VN
W2 = <I> = R + h' (2.15.3)

W3 = (). + U) sin <I> = (RV~ h + U cos <I> ) tan <I> ,

where U stands for the angular velocity of the earth. Equations (2) and
(3) determine the translational motion. The point coordinates in the axes
Oxyz are equal to (0,0, h) and thus the components of its relative velocity
along these axes are as follows

Vrl = 0, Vr 2 = 0, Vr3 = h= Vh· (2.15.4)

The absolute velocity of the point, i.e. the velocity relative to the axes
keeping constant directions relative to the unmovable stars, are given by
92 2. Rigid body kinematics - basic knowledge

the following expressions for its projections on these axes

Val = Vi + W2h = VN, }


Va2 = "!2 - wlh = -Va - (R + h) U cos <I> , (2.15.5)
Va3 = h.
We proceed now to calculation of acceleration. Acceleration w of the origin
oof the axes Oxyz is given by

Applying eq. (7.5) yields

v= ~ h VN). + (R ~ h Va + RU cos <I> ) i~ +


r
w = [( R W3]

[- (R ~ h Va + RU cos <I> + R ~ h VN W3] i~ -


[R~ h VN W2 + (R~ h Va + RUCOS<I» Wi] i~.
The projections are given by
2
R. R Va
Wi = -R
+
h VN -
(R+h)
2VNVh + R ( - R
+
h + U cos <I> ) tan <I> ,
R. R VN . RVNVa
W2 = - R hva + 2 VaVh + 2RU - R h sm <I> + 2 tan <I> ,
+ (R+h) + (R+h)
2
R 2 Va
W3 = -
(
2 V N -R - R h + U cos <I> )
(R+h) +
(2.15.6)
To construct an expression for the translational acceleration it is also
necessary to calculate the following vector

ex r' + w x (w x r') = e x r' + ww· r' - w 2 r',


where r' = hi~. Projections of this vector on the axes are

(€2 + W1W3) h, (-€l + W2W3) h, - (wi + w~) h,


respectively. The projections of the angular acceleration vector are equal to
the time derivatives of the corresponding projections of the angular velocity,
i.e.

(2.15.7)
2.16 Body rolling on a fixed plane 93

Projections of the relative acceleration on axes Oxyz are equal to (0,0, Vh)
and the Coriolis acceleration is as follows

Now we can determine the absolute acceleration. Its projections are

Wal WI + (E2 + WIW3) h + 2W2Vh,


Wa2 W2 + (-EI + W2W3) h - 2WIVh,
Wa3 W3- (wi+w~)h+Vh.

By means of the above formulae we obtain

.
Wal =VN+ tan <I> + 2Vo U'
VNVh + vbR+h
R+h "'" (R + h) U2 cos ""'.
sm,*,+ ,*,sm "'"
'*',
. VOVh VNVO .
Wa2=-VO- R+h + R+htan<I>+2U(vNsm<I>-vhcoS<P),
v 2 +v 2
Wa3 = Vh - ~ + hO - 2Uvo cos <P - (R + h) U 2 cos 2 <P.
(2.15.8)

Another form of these equalities is given in [69].

2.16 Body rolling on a fixed plane


As another example of constructing equations for non-holonomic constraints,
calculation of three-index symbols and application of the results of Sec.
2.13, we consider the case of a body rolling without slipping on a fixed
plane. The body surface S is assumed to have a tangent plane at any
point.
Let a system of axes Ox' y' z' be fixed in the body and have the pole (the
origin) at point 0 of the body. We introduce the Gaussian coordinates ql
and q2 on surface S, then the position vector p with the origin at the pole
o is considered as a function of ql and q2.
Following [91] we will describe the position of the body with the help of
five generalised coordinates. Two of them are the Gaussian coordinates ql
and q2 of point M at which the surface S makes contact with the plane.
Another two coordinates are the Cartesian coordinates x and y of this
point in the plane relative to the fixed coordinate system Oxyz, axis z
being normal to the plane. The four functions of time

prescribe curve L' which is a locus of the contact point M on the surface S
and curve L which is a locus of the contact point M on the plane, see Fig.
94 2. Rigid body kinematics - basic knowledge

~ Y L

FIGURE 2.14.

2.14. Curves L' and L have a common tangent at point M characterised


by the tangent unit vector r. The body position is prescribed when we
introduce the fifth generalised coordinate which is angle {} between the
vector r and an axis fixed in the plane, for example axis Ox.
Following Sec. B.7 we introduce the covariant basis vectors PI and P2
on surface S and the unit vector m of the normal to the surface S. The
direction of the latter vector and axis Oz coincide. Then, denoting the
angular velocity vector of the body (the axes Ox'y'z') by wand applying
relationship (13.3), we have
• *
m=m+w x m=O, (2.16.1)

as m is invariable in space. In the latter equation m* is the time derivative


of m in axes Ox'y'z'. In view of (B.7.24)
* - maq'a -_ - baf3Pf3·a
m- q , (2.16.2)

where pf3 are the contravariant basis vectors and ba f3 are coefficients of the
second quadratic form of the surface.
Projection of vector w on the normal m is denoted in sequel as n = w . m.
The vector product of m and eq. (1) is

mx ~ +m x (w x m) = 0,
which yields

But

(2.16.3)
2.16 Body rolling on a fixed plane 95

Here aa{3 denote the coefficients of the first quadratic form, i.e. the covariant
components of the metric tensor of the surface, a~ stands for its mixed
components and lal denotes the discriminant of the first quadratic form.
Now we obtain

(2.16.4)

which gives an expression for vector w in terms of the quantities determined


by the surface geometry and [2.
Projecting the latter equation on the axes fixed in the body we have

In order to understand the geometrical meaning of parameter [2 we notice


that vector 7- can be represented in two ways. First it is the derivative of
the unit tangent vector to the curve L and for this reason

7- =~m x 7. (2.16.6)

Alternatively,

d

7 =7
*
+w x 7 = m ba{3q.a d;;
q{3
+ Po: k*"·a + w x 7. (2.16.7)

Here relationships (B.8.3) and (B.8.6) are used, k*" denotes contravariant
components of the vector of geodesic curvature and da is the arc element
of curve L'. Substituting expression (B.8.2) for 7 we obtain

WX7

As one may expect, the term corresponding to the vector of the normal
curvature in eq. (7) vanishes. Inserting the latter expression for W x 7 into
96 2. Rigid body kinematics - basic knowledge

eq. (7) and comparing the result with eq. (6) we arrive at the following
equation

{j = n + Pc< . (m x r) k*" a. (2.16.8)

In addition we have

Pc< . (m x r)

and

Thus we have

(2.16.9)

It follows from this equation that for given ql (t) ,q2 (t) and n (t) angle
{) is determined by a quadrature, and then x and yare determined by
quadratures from the following relationships

dx = cos {)da, dy = sin {)dCT (2.16.10)

since in the case of rolling without slipping the arc element da is the same
for curves Land L'. This problem is solved in Sec. 8.3.
Let v denote the velocity vector of the pole 0 and Vs its projections on
the axes fixed in the body. As the velocity of contact point M of the body
with the plane is equal to zero, we arrive at the following three equations

Xl = VI + W2 Z ' - W3Y' = 0, }
X2 = V2 + W3X' - WlZ' = 0, (2.16.11)
X3 = V3 + WlY' - W2X' = 0.

Next we introduce the notation

(2.16.12)

·1 ·2
X5 = q, X6 = q , (2.16.13)

where WI, W2, W3 are determined by means of eq. (5). Quantities Xs are taken
as quasi-velocities, three of them vanish by virtue of equations (11) for non-
holonomic constraints. The corresponding variations of quasi-coordinates
2.16 Body rolling on a fixed plane 97

are denoted by 87r s' To shorten the equations we introduce vectors X and 87r
whose projection on the axes fixed in the body are Xs and 87r s (s = 1,2,3).
Then

X = v +w x p, 87r = 8r + 0 x p, (2.16.14)

and formulae (10.2) and (10.6) can be written in vector form as

o* -8w = 0 x w (8r)* - 8v = 8r x w - v x O. (2.16.15)

Here an asterisk designates the time derivative of a vector relative to axes


Ox' y' z' fixed in the body. This implies that projections of vectors 0* , (87r)*
etc. are equal to the time derivative of the projections of these vector::; on
the above axes. Thus we obtain

(87r)* - 8X = 8r x w - v x 0+ (0 x w) XP + Ox p -w x 8p. (2.16.16)

Replacing now 8r and v by their values due to eq. (14) and taking into
p
account that and 8p are equal to Pat'> and Pa8qa, respectively, we arrive
at the following expression

(2.16.17)

where the following terms vanish

- (0 x p) xw+ (w x p) x 0+ (0 x w) XP = O.

What remains is to replace wand 0 on the right hand side of eq. (17) by
their expressions due to eq. (4). The result is as follows

(2.16.18)

where vectors Ia are given by eq. (4). Introducing the notation

(2.16.19)

we obtain

(2.16.20)

The right hand side contains only quasi-velocities and the corresponding
variations of quasi-coordinates. By projecting vector relationship (20) on
axi::; Ox' we obtain the three-index symbols 1~(3 which are coefficients of
the products Xa87r(3, where (l = X5 and (? = X6 due to eq. (13). The tables
of the three-index symbols having superscripts 1,2 and 3 are given below.
98 2. Rigid body kinematics - basic knowledge

I cx""'(3 I 1 I 2 3 4 5 6
1 I0 0 0 0 0
2 0 -m3 -h3 -l23
3 m2 h2 lz2
4 -k11 -k21
5 0

Table of '~/3

I cx""'(3 I 1 I 2 3 4 5 6
1 I0 0 m3 l13 l23
2 0 0 0 0
3 -m1 -l11 -l21
4 -k12 -k22
5 0

Table of ';/3

I cx""'{3 I 1 I 2 3 4 5 6
1 I0 0 -m2 -h2 -l22
2 0 m1 l11 l21
3 0 0 0
4 - k 13 - k 23
5 0

Table of ';/3
Here kas and las are interpreted as projections of vectors on axes Ox' y' z',
for example

The difference (87r4r - 8X4 can be written as follows

(m. Or - 8 (m· w) = m· (O-8w)


m· (w x 0) = (m x w) ·0
2.16 Body rolling on a fixed plane 99

as m = 0 and 8m = o. Substituting expression (10.10) for difference 9-8w


and expressions (18) for (J and w into the latter equation we obtain

(2.16.21)

Using projections of vectors la and 113 and recalling the definition of m we


find that

hence

I: )- I: bll b22 - bf2 (.1 I: 2 ·2I: 1)


(u7r 4 - uX4 = .Jjai q uq - q uq . (2.16.22)

Due to eq. (20) the non-zero three-index symbols having superscript 4 are
only

(2.16.23)

All three-index symbols having superscripts 5 and 6 are identically equal


to zero since expressions X5 = q1 and X6 = q2 are integrable.
We determine the angular acceleration vector by differentiating eq. (18)
for the angular velocity vector with respect to time

By virtue of eqs. (B.7.24) and (B.7.19) we find

rna = -b~Pi3 = -bai3pi3 , lai3 = (P2 a~i3 ~ - P1 a~i3 ~) +


~ [ba1 ({ 2~ } P, + b2i3m) - ba2 ({ 1~ } P, + b1i3 m)] ,

or
100 2. Rigid body kinematics - basic knowledge

FIGURE 2.15.

Noticing that

we obtain

As an example we consider the case of a rigid body bounded by a surface


of revolution. Axis 0 z' is directed along the axis of symmetry of the surface,
then

x' = v (8) COScp, y' = v (8) sincp, z' = Z' (8),


see Fig. 2.15.
The Gaussian coordinates are the arc 8 measured along the meridian
arc from the point of intersection of the meridian and axis 0 z' and the
azimuthal angle cpo The distance between the point and axis Oz' is denoted
as v. The unit vector of the normal m to the surface is also normal to the
2.16 Body rolling on a fixed plane 101

meridian. If a denotes the angle between the normal m and axis Oz' then
it is easy to see that

dv dz'
ds = cos a, ds = -sina.
We obtain the following expressions for the projections of vectors PI, P2
and m on the axes Ox' y' z'

Ox' Oy' OZ' II

PI cosacoscp I cosasincp I -sina II

P2
m
-vsincp
sin a cos cp
v cos cp
sin a sin cp °
cos a
II"

II
Then we find

au PI . PI = 1, = PI . P2 = 0, a22 = P2 . P2 =
aI2 v 2,
8m da
v, bu = - - . PI = - - = -k,
8s ds
8m 8m.
- 8s ·P2 =0, b22 = - 8cp ·P2 = -vsma,

where k denotes curvature of the meridian. The non-zero Christoffel sym-


bols of the second kind for the surface of revolution are

{ 2 } ={ 2 } = cosa
12 21 v '

whereas the components of the vector of the geodesic curvature for the
curve s = s(a) and cp = cp(a) are given by

By virtue of eq. (9) we obtain

(2.16.25)

Formulae for projections of the angular velocity vector on the axes Ox' y' z'
take the form

(0 + cpcos a) sina cos cp + ks sin cp, }


(0 + cp cos a) sin a sin cp - ks cos cp, (2.16.26)
Ocosa - cpsin 2 a.
102 2. Rigid body kinematics - basic knowledge

tz ;c, !I'

~
r

FIGURE 2.16.

2.17 Composition of motions of a rigid body


The problem of composition of motions of a rigid body is a generalisation
of the theory of the relative motion explained in Sec. 2.14. Not the particle
M but the rigid body S is now the object which makes a prescribed motion
relative to the system of moving axes Ox' y' z'. Relative motion of particle M
was described in Sec. 2.14 by means of given dependences of the coordinates
on time, i.e. x'(t), y' (t) and z'(t). However the relative motion of the body
must be described by the motion of its pole, that is point M in Fig. 2.16, and
the relative velocity Wr relative to the axis system Ox' y' z', i.e. the relative
angular velocity. The system of axes M x*y* z* is fixed in the rigid body
S. The motion of a generic point N characterised by the position vector
p =MN with respect to the basic system Oxyz is sought. The angular
velocity of the rigid body which is mentally bound to the axes Ox' y' z' is
denoted now as We and is called the translational angular velocity. It was
denoted as W in Sec. 2.14.
One can immediately apply the theorem of velocity composition derived
in Sec. 2.14 when one forgets for the moment about the rigid body and
focuses attention on point N. In this case the translational velocity Ve is
understood to be the velocity of the coincident point, i.e. the point of a
fictitious body mentally bound to the system Ox' y' z' which coincides with
point N instantaneously

Ve = Vo + We X (r' + p) ,
where r' = oM.
By virtue of eq. (7.8) for the velocity of a point in the rigid body, the
relative velocity of point N with respect to the system Ox' y' z' is the sum
of the velocity of the pole M (denoted earlier as v r) and the rotational
velocity Wr x p. Thus, the absolute velocity v A of point N is equal to

VA = Vo + We X (r' + p) + Vr + Wr X P
2.17 Composition of motions of a rigid body 103

or

VA = Vo + We X r' + Vr + (We + Wr) X p. (2.17.1)

The first three terms represent the absolute velocity of point M denoted
by Va. Then

(2.17.2)

In particular, let points 0,0 and M coincide. Then Va = 0 and the velocity
of point N is as follows

(2.17.3)

But in this case, forgetting for the time being about the existence of axes
Ox' y' z', we would write that the velocity of point N is given by

VA = WA X p, (2.17.4)

where WA denotes the angular velocity with respect to axes Oxyz which
implies that W A is the absolute angular velocity. Comparing eqs. (3) and
(4) we conclude that because of arbitrariness of the position vector p we
have

(2.17.5)

The result obtained is the theorem of addition of angular velocities. More-


over, if we set p = i; in eq. (3), i; being unit vector of the s - th axis of
system M x*y* z* we obtain

(2.17.6)

which also follows from eqs. (5) and (7.5).


Vector We is naturally prescribed by its projections on axes Ox' y' z' whilst
vector Wr is prescribed by projections on axes M x*y* z*. For this reason,
the angular velocity vector is
3

cA WA = L (wesi~ + wwi;t
s=l
3
L (wesi~ + wrs*i;) + We X We + (We + Wr) X Wr,
s=l

or

(2.17.7)
104 2. Rigid body kinematics - basic knowledge

where e r denotes the vector whose projections on axes M x*y* z* are equal
to the time derivatives of the projections of the relative angular velocity
Wr on the same axes.
Taking now the derivative of expression (2) with respect to time we deter-
mine a formula for the absolute acceleration W A of point N. Differentiating
Va was performed in Sec. 2.14 and gave formula (14.13) for absolute accel-
eration Wa of point N. In this formula one must replace wand e by We and
ee, respectively. The time-derivative of the second term in eq. (2) remains
to be taken. Using eqs. (7) and (3) we obtain

[(We + wr) X pt = (ee + e r + We X wr)xp+ (We + Wr)X[(W e + Wr) X p]


= ee X p + We X (We X p) + (We X Wr ) X P + We X (W r X p)
+ Wr X (We X p) + e r X p + Wr X (W r X p) .
Noticing that

we can write the previous expression in the form

[(WeXWr)Xpt = eexp+wex(wexp)+erxp+
Wr X (W r X p) + 2we X (W r X p) .
The final expression for WA takes the following form

{Wo+e e X (r'+p)+we X [we X (r'+p)]) (2.17.8)


X (wr X pH + {2we X [v r + (w r X p)]).
+ {wr + er X p + wr
The quantity

WE = Wo + ee X (r' + p) + We X [We X (r' + p)] (2.17.9)

comprises translating acceleration, i.e. the acceleration of point N which is


mentally rigidly bound to axes Ox' y' z'.
The second group of components, namely

(2.17.10)

is the relative acceleration of point N with respect to axes Ox' y' z' and is
calculated by means of the formula for acceleration of a point in the rigid
body (11.1). The first term in eq. (10) presents acceleration ofthe pole M,
whereas the second and the third ones comprise rotational and centripetal
accelerations, respectively, both calculated under the assumption that the
axes Ox'y' z' are fixed. The relative velocity of point N with respect to axes
Ox' y' z' is equal to

VR = Vr + Wr X p. (2.17.11)
2.18 Motion of the natural trihedron of a spatial curve 105

FIGURE 2.17.

Because of this, the third component

(2.17.12)

represents the Coriolis acceleration of point N. This yields the theorem of


composition of accelerations

WA =WE+WR+WCor' (2.17.13)

Expression (8) is very useful as it explains comprehensively the meaning of


each component in eq. (13).

2.18 Motion of the natural trihedron of a spatial


curve
The position of point M of a spatial curve is given in the system of fixed
axes Oxyz by the position vector r which is considered as a prescribed
function of a curvilinear coordinate a counted from this curve from the
origin Mo.
The unit vector T that is tangent to the curve is given by the following
equality
dr
T=-. (2.18.1)
da
Let Tl = T+dT denote the tangent vector at point M' which is infinites-
imally close to point M. When we place vector Tl to point M as shown in
Fig. 2.17 then vectors T and Tl determine the tangent contact plane II at
point M. An infinitesimal vector dT is perpendicular to T and directed to
the concave side of the curve. Its value IdTI = c defines an infinitesimal
angle between T and Tl and the ratio
106 2. Rigid body kinematics - basic knowledge

determines the curvature of the curve at point M, p being the radius of


curvature at this point. Therefore, the vector
dT
n=p- (2.18.2)
dO'
is a unit vector which lies in the tangent plane and is coincident with the
normal toward the concave side of the curve. For this reason n is referred
to as the unit vector of the principal normal to the curve. A unit vector of
binormal is constructed using the rule b = T X n, see Fig. 2.17.
Thus, an orthogonal trihedron T, n, b is determined at any point M of
the curve C. Let us consider a certain point N fixed in this trihedron. We
denote the position vector MN of this point relative to the trihedron origin
as r' and its projections on the axes of the natural trihedron as aI,a2,a3.
Let
• dr. .
Vo = r=-O'=TO'
dO'
designate the velocity of the trihedron origin. The velocity of point N is
then

Alternatively, due to the formula for the velocity of a point in the rigid
body

v Vo+0 x r' (2.18.4)


Ta + T (02a3 - 03a2) + n (03a1 - 01a3) + b (01a2 - 02a1) ,

where 0 denotes the angular velocity vector of the natural trihedron whilst
moving along the curve.
As a1,a2,a3 are arbitrary, a comparison of relationships (3) and (4) en-
ables the following three equations to be determined

(2.18.5)

It follows from the first equation and (2) that

The value of 0 1 is determined from the third equation in (5)

n
H1 dO' . n = -a. ( T x dn)
= -a. db . ( n x dn)
dO' . n =O'T· dO' . (2.18.6)
2.18 Motion of the natural trihedron of a spatial curve 107

It is easy to obtain this result by differentiating b = T X n and accounting


for eq. (2) which yields
db dn
-=TX-.
dO" dO"
The value
x' (0") y' (0") Z' (0")
x" (0") y" (0") Zll (0")
x"' (0") ylll (0") Zll' (0")

(2.18.7)

is termed the twist of the curve. As follows from eq. (2), the curvature of
the curve is given by the following equality

-1 =
P
I-dT I =
dO"
V
[XII (0")] 2 + [y" (0")] 2 + [Zll (0")] 2 . (2.18.8)

Taking into account eqs. (5)-(7) we come to the well-known Frenet formulae
dT n dn b T db n
(2.18.9)
dO" p' dO" T p' dO" T
and the expression for the angular velocity of the natural trihedron

n =& (~+;). (2.18.10)

The quantity in the parenthesis is named the Darboux vector. Let us


notice that the third formula in (9) says that the move from point M to
the infinitesimally close point M' results in a rotation of the tangent plane
about the tangent T of
dO"
Idbl = TJ = TTl'
thus the binormal b gets (for T > 0) an increment db which is opposite in
direction to n.
We consider now the motion of a rigid body, for instance an airplane,
one point of which (the pole M) moves along the curve C. Adopting the
notation of Sec. 2.4 we introduce a system of velocity axes M x*y* z* whose
position with respect to the fixed axes O~TJ( is described by angles A, /-l, lJ.
The directions of axes M x* and T coincide, thus plane M y* z* and the
normal plane of the trajectory coincide also. Figure 2.18 shows the rotation
through angle X about T that makes the unit vectors nand b coincident
with axes M y* and M z*. Then, designating the angular velocity vector of
the velocity axes as w', we have

(2.18.11)
108 2. Rigid body kinematics - basic knowledge

b n"

FIGURE 2.18.

i; being the unit vectors of the velocity axes. Alternatively, the definitions
of angles >., f-I" v yields the following expressions for the projections w~* of
vector w' on the velocity axes i;

wi* = '\sinf-l, + it, }


w;* = ,\ cos f-I, cos v + jJ, sin v, (2.18.12)
w~* = -'\cosf-l,sinv+jJ,cosv.
We obtain three equations

~+X='\sinf-l,+it, (2.18.13)

~ sin X = ,\ cos f-I, cos v + jJ, sin v , (2.18.14)


P

~ cos X = -,\ cos f-I, sin v + jJ, cos v. (2.18.15)


P
From the latter two equations we find

(2.18.16)

cos (X - v) = P# =
a JjJ,2 + jJ,,\2 cos2 f-I,
')
(2.18.17)
. ( ) '\COSf-l, '\COSf-l,
---,.======
sm X - v = P--.-
a
=
JjJ,2 +,\2 cos 2 f-I,
2.18 Motion of the natural trihedron of a spatial curve 109

Differentiating the first equation in (17), making use of the second equa-
tion and the expression for X - v from (13), we arrive at the following
relationship

iJ .. jL2,X sin fL + (,XjL - ).jL) cos fL


- = A sm fL + .2 . (2.18.18)
T jL2 + A cos 2 fL
Formulae (16) and (18) determine the motion of the natural trihedron
for prescribed functions of time A (t) and fL(t). Given the velocity axes,
the directions of axes of the natural trihedron are found by formulae (17).
Sections 2.4 and 2.5 are devoted to the construction of the velocity axes
when the axes fixed in the body are prescribed.
3
Theory of finite rotations of rigid
bodies

3.1 Rodrigues formula and the vector of finite


rotation
A rigid body having a fixed point 0 is subject to rotation through an angle
X about an axis whose direction is given by unit vector e. The direction of e
is chosen in such a way that watching from the end of vector e one observes
the rotation through a positive angle ::;: 180 0 , that is counterclockwise for
a right-handed coordinate system.
Let us consider the position vector oM = p before the rotation. After
----),

the rotation it takes position OM' = p', so that the vector

~
MM' =p'_p (3.1.1)

represents the displacement of point M caused by the body rotation. This


displacement is required to be expressed in terms of the rotation parame-
ters, i.e. angle X and axis e, and vector p.
It is clear that either vector p and p' is a generator of a cone whose
axis coincides with unit vector e as shown in Fig. 3.1. The component of
vector p along the rotation axis, i.e. the vector ee· p, does not change
under rotation, thus

e· p = e· p'. (3.1.2)
112 3. Theory of finite rotations of rigid bodies

o
FIGURE 3.1.

What remains is to observe a change in the component perpendicular to


the axis. Before the turn it is

Po = P- ee· p, (3.1.3)

whilst after the turn it is

p~ = p' - ee . p' = p' - ee . p. (3.1.4)

Figure 3.2 provides a view from the end of vector e showing that

Observing that vector 8M) has the magnitude 0 1 8 tan ~ and the direction
coinciding with that of the vector product

we obtain

, 1(
Po='2 Po + Po') + '12 e x (
Po + Po') tan '2X
or
, ,X X
Po - e x Po tan '2 = Po + e x Po tan '2'

Replacing in the latter equation p~ and Po respectively by p' and p due


to eqs. (3) and (4), Rodrigues formula is produced

p' - e x p' tan ~ = p + e x p tan ~. (3.1.5)


3.1 Rodrigues formula and the vector of finite rotation 113

0,

'-------~------~~,

FIGURE 3.2.

This equation remains to be resolved into p. To this end, we calculate the


vector product of both sides of eq. (5). Taking into account that
ex (e x pi) = ee· p - p'

we obtain

p' tan ~ +e x pi = - p tan ~ +e x p + 2ee . p tan ~. (3.1.6)

Eliminating vector e x p' from eqs. (5) and (6) we have

pi _ 1 [p (1 _ tan 2 ~) + 2e x ptan ~ + 2ee. ptan 2 ~] .


- 1 + tan 2 ~
(3.1.7)
This formula can be recast as follows

p' = P + 1 [e tan ~ x p + e tan ~ (e tan ~ . p) - ptan 2 ~] .


1 + tan 2 ~
(3.1.8)
The two latter terms comprise the following product

e tan ~ x (e tan ~ x p) ,
thus formula (7) can be rewritten in the form

2etan ~ X
pi = P + 2X x (p + e tan - x p) . (3.1.9)
1 + tan 2 _ 2
2
We introduce into consideration vector (J whose direction coincides with
unit vector e of the rotation axis and whose magnitude is

e=I(JI=2tan~. (3.1.10)
114 3. Theory of finite rotations of rigid bodies

Then we obtain the final form of formula (9) suggested by Rodrigues

,
P = P+ 1 1 2 () X
1 + 48
(1
+ -() )
P
2
x p . (3.1.11)

The introduced vector

() = 2etan ~ (3.1.12)

is termed the vector of finite rotation. In what follows the word "finite"
will be omitted unless this leads to confusion. Any manipulation of vector
algebra is applicable to this vector. For example, expressing e in terms of
the unit vectors is of a Cartesian basis Oxyz

e = i1 coso: + i2 cos (3 + h cOS'Y,


0:, (3 and 'Y being angles between e and is, we are entitled to write
(3.1.13)

where the quantities


- X - X - X
81 = 2 tan "2 cos 0:, 82 = 2tan"2 cos (3, 83 = 2 tan "2 cos'y (3.1.14)

comprise projections of vector () on axes Oxyz. However it would be erro-


neous to treat notation (13) as a statement that finite rotation described by
vector () can be achieved by means of three rotations about coordinate axes
through angles Fh, B2 and B3 . This would be the case when finite rotations
would obey the law of vector addition as forces or velocities do. However,
in reality, the law of commutation of rotation is more complex, see Sec. 3.3
for detail.

3.2 Parameters of Rodrigues and Hamilton


Let us consider the vector p that is bound to the moving axes Ox' y' z' that
coincided with the fixed axes Oxyz before the rotation took place. Vector
() performs the rotation of axes Oxyz into Ox' y' z', thus taking p = is we
will have p'= i~ and by virtue of (1.11)

. ,.
Is = Is + 11 2 () X
(.Is + -1 () .)
X Is . (3.2.1)
1 + 48 2
Rotation -() makes the transition of axes Ox' y' z' into Oxyz. In this case
p = i~ and p'= is and we obtain a relationship which is inverse to (1)

. .,
Is = Is -
11 2 () X
(.,
Is -
1 ()
-
., )
X Is . (3.2.2)
1 + 48 2
3.2 Parameters of Rodrigues and Hamilton 115

It follows from these formulae, as well as from eq. (1.2) that

(3.2.3)

i.e. projections of the rotation vector on the corresponding axes of coordi-


nate systems Oxyz and Ox' y' z' are equal to each other

Os = Os (8 = 1,2,3) . (3.2.4)

Expression (1.13) can be written in either of the following two forms

(3.2.5)

The formulae of the theory of finite rotation will look more symmetrical
if, instead of projections of the rotation vector, one uses the quantities
proportional to them, namely

(3.2.6)

The proportionality factor is chosen under the following condition

(3.2.7)

Its geometrical meaning is easy to understand. Due to eq. (1.10) we have

1 2 2 X 1
1 + -40 = 1 + tan -2 = --2-'
cos X

On the other hand, it follows from eqs. (6) and (7) that

Comparing these equations we have

(3.2.8)

Thus, the rotation vector () is determined by the four Rodrigues-Hamilton's


parameters
\ . X \ . X X
/\1 = cos a SIn 2' >'2 = cos (3 sin ~, /\3 = cos 'Y SIn 2 ' >'0 = cos 2'
(3.2.9)

subject to a single condition (7). Here a, (3, 'Yare angles of rotation relative
to the basis axes (coinciding in coordinate systems Oxyz and Ox' y' z') and
116 3. Theory of finite rotations of rigid bodies

X is the rotation angle. An expression for () and formulae (1) and (2) take
the form

(3.2.10)

(3.2.11)

3
.Is = ./
Is -
2~'·/
~ "'tIt x (3.2.12)
t=l
Recalling the definition of the Levi-Civita symbols and using the following
equality

we can transform relationships (11) and (12) to the form

(3.2.13)

is = i~ - 2 (>.0 t Etsq>'ti~ t >'qi~ + i~ t >.~) .


t=l
- As
q=l q=l
(3.2.14)

These yield expressions for the direction cosines of angles between the axes
of these coordinate systems

(3.2.15)

(3.2.16)

Recalling the values of the Levi-Civita symbols and making use of relation-
ship (7) we obtain the table of direction cosines. It is constructed row-wise
by means of eq. (15) and column-wise byeq. (16).

II II x y
>.2o + >.21 _ A22 _ >.23 2 (>'OA3 + A1>'2)
I z II
I Xl II I +
2 (->'0>'2 >'1>'3) II
II yl II 2 (-AO>'3 + >'2>'d I >'6 + A~ - >.~ - >.i I 2 (>'0>'1 + A2>'3) II
II zI II 2 (>'0>'2 + >'3A1) I 2 (-AO>'l + >'3 A2) I >.2+>.2_>.2_A2
o 3 1 2 II
3.3 Composition of finite rotations 117

Table 1 of cosines in terms of the Rodrigues-Hamilton parameters

Using eq. (15) we can also introduce a notational shorthand for the co-
ordinate transformation
3 3 3
Xs = (A6 - Ai - A~ - A~) XS + 2Ao L L EtskAtXk + 2As L AkXk,
k=lt=l k=l
(3.2.17)

3 3 3
Xs = (A6 - Ai - A~ - A~) X~ - 2Ao L L EtskAtX~ + 2As L AkX~.
k=lt=l k=l
(3.2.18)

Let us draw your attention to the fact that the coordinate transformation
(18) is the result of projecting the Rodrigues formula on axes Oxyz. The
coordinate transformation (17) is obtained when projecting the relationship

P = P, - g x (p' - -()
1 x P') (3.2.19)
1 + 192
4
2

on axes Ox' y' z'. Equation (19) is an inverse form of the Rodrigues formula
and expresses that rotation -() makes vector p' coincident with p, as well
as axes Ox' y' z' coincident with Oxyz.

3.3 Composition of finite rotations


Let a rigid body, having an immovable point 0, be subjected to rotation
()l, i.e. it rotates through an angle Xl about the axis prescribed by the
unit vector el. Afterwards the body is subjected to a second rotation ()2
characterised by an angle X2 and the unit vector e2. The vectors el and
e2 are assumed not to be related to the body. The angle (less than 180°)
between these vectors is denoted by <I> , then

(3.3.1)

The Euler-Chasles theorem states that any rotation of the rigid body,
with one point fixed, is always equivalent to a single rotation about a line
through this point. This rotation, described by vector (), angle X and axis
e, is required to be expressed in terms of vectors ()l and ()2.
We construct a spherical triangle ABC on the unit sphere about 0 as
shown in Fig. 3.3a. Its vertices coincide with the ends of the unit vectors
el, e2 and e3. The vertex C corresponds to the rotation sequence ()1,()2.
This vertex can be constructed as the point of intersection of arcs AC and
118 3. Theory of finite rotations of rigid bodies

-4

o
FIGURE 3.3.

BC with angles ~Xl and ~X2 relative to arc AB as shown in Fig. 3.3.
Indeed, rotation (h carries point C to C' whilst rotation (}2 brings C' back
to C because the spherical triangles ABC and ABC' are equal. From this
construction one can see that rotations are not commutative since rotation
(}2 followed by rotation (h does not result in point C but in G'.
The direction of the axis of the resultant rotation is now defined. What
is left is to determine angle X. To do this we now turn to Fig. 3.3b. Point
A remaining fixed at the first rotation takes position A' after the second
rotation. This is also the position of this point at the resultant rotation ()
through angle X. Thus the angle at vertex C of the spherical triangle ABC
is equal to ~(27r - X) = 7r - ~X.
Now we derive the formulae corresponding to the above geometric con-
struction. We will use formulae of spherical trigonometry

cose = cosacosb + sinasinbcosC, (3.3.2)

sin a sin b sin e


(3.3.3)
sin A sinB sinG'

cos C = - cos A cos B + sin A sin B cos e, (3.3.4)

see Fig. 3.3 for notation. Equation (4) for cos C is less well-known than the
others. It can be easily derived by applying eq. (2) to the polar spherical
triangle A* B*C* whose vertices lie at the poles of great circles forming the
sides of triangle ABC.
Applying eq. (4) to the spherical triangle ABC shown in Fig. 3.3 we
obtain

cos -X = cos -Xl cos -X2 - sm


. -Xl sm
. -X2 cos '¥
;J,.
(3.3.5)
2 2 2 2 2 '
3.3 Composition of finite rotations 119

which by virtue of eqs. (1) and (2) can also be written in the form

X
cos "2 =
Xl
cos 2 X2 ( 1 -"4
cos 2 1 () ())
I· 2 . (3.3.6)

We derived expression for the angle of the resultant rotation. We express


vector e in terms of the non-coplanar vectors el, e2 and e2 x el

(3.3.7)

where the coefficients are to be determined. Coefficient r is positive as


vector e (see Fig. 3.3) deviates from the plane of vectors el and e2 to the
side of e2 x el. The case r < 0 corresponds to the case in which rotation
(}2 precedes rotation (}l.
Since e is a unit vector, eq. (7) yields

or

r = -._1_
sm
V1 - a 2 - (32 - 2a(3cos<l>.
<I>
(3.3.8)

In order to obtain a and (3 without solving any equation we first calculate


the scalar product of eq. (7) and e2 x (e2 x el), and second the scalar
product of eq. (7) and el x (e2 x ed. Performing these operations we obtain

Then by means of Fig. 3.3a and formulae (1) and (2) we obtain

e . e2el . e2 - e . el
cos a cos <I> - cos b = - sin a sin <I> cos ~2 ,

e·e2 -e·elel·e2
cos a - cos b cos <I> = sin b sin <I> cos ~l ,
120 3. Theory of finite rotations of rigid bodies

Now making use of eq. (3) we find

----+
. Xl . X2
sin a X2 sm 2 X2 sinb X sm- X
a = -:---<I> cos -2 = --X- cos -2 ' f3 = -.- cos --1 = cos --1.
sm sin- sm <I> 2 sin _ 2
2 2
(3.3.9)
By means of eq. (3) we obtain

'Y = 1 (sin2 ~ _ sin2 Xl cos2 X2 _


sin<I>sin~ 2 2 2
2
X cos 2 --1
sin 2 --1 X - 2 sin --1
X sin --1
X cos --1
X cos --1
X )1/2 .
2 2 2 2 2 2

Replacing here sin 2 ~ = 1- cos 2 ~ by its expression due to eq. (5) we find
after simple algebraic manipulations

sin Xl sin X2
'"V- 2 2 (3.3.10)
1- • X
sm 2

Inserting expressions (9) and (10) into eq. (7) we have

e = -:---x
sm-
1 ( . Xl X2 . X2 Xl . Xl . X2 )
el sm -2 cos -2 + e2 sm -2 cos -2 + e2 x el sm - sm -
2 2
,
2
(3.3.11)
which by means of the definition for the rotation vector can be cast in the
form
cos Xl cos X2 1
() = 2 X 2 ((}l + (}2 + -(}2 x (}l) . (3.3.12)
cos- 2
2
By virtue of eq. (6) we arrive at the formula expressing the rule of compo-
sition of finite rotations

(3.3.13)

The presence of the term (}2 x (}l indicates that finite rotations are not
commutative. That is, the resultant rotation of (}2 followed by (}l is given
by the following formula

(3.3.14)
3.3 Composition of finite rotations 121

In the particular case of two rotations about the same axis, eqs. (13) and
(14) express theorem on tangent of a sum. For

(h = 2etan ~1, (h = 2etan ~2

we have

0= Of = ~e X (tan Xl + tan X2) = 2etan Xl + X2.


1 - tan ---.!. tan --.2 2 2 2
2 2
(3.3.15)

It is easy to obtain expressions for the Rodrigues-Hamilton parameters


of the resultant rotation (denoted by vo, VI, V2, V3) in terms of tho:;e pa-
rameters AO, AI, A2, A3 for the first rotation and f.Lo, f.L1' f.L2' f.L3 for the second
rotation. First by means of eqs. (2.8) and (13) we calculate the following
value

Since

we obtain

1 (1 + iBn (1 + iB~) 1
v6 - (1 - i 0 1 . ( 2 )2

and thus we can adopt that


3

Vo = Aof.Lo - L Asf.Ls· (3.3.16)


s=l

Now by means of eq. (13) we obtain

__
\
A. . :.O'-: ~"-O_ _ (~s + f.Ls + -A_1-
'" \ 0 f.Lo
tt
of.Lo r=l t=l
ErtSf.LrAt)
/lof.Lo - L.J /lsf.Ls
s=l
or
3 3

Vs = Asf.Lo + f.LsAo + L L Ertsf.Lr At . (3.3.17)


r=lt=l
122 3. Theory of finite rotations of rigid bodies

From eqs. (16) and (17) we obtain the following system of formulae

110 = Ao/Lo - A1/L1 - A2/L2 - A3/L3' }


111 = A1/LO + Ao/L1 + A3/L2 - A2/L3' (3.3.18)
112 = A2/LO + Ao/L2 + A1/L3 - A3/L1,
113 = A3/LO + Ao/L3 + A2/L1 - A1/L2'

It is worthwhile noting that these expressions present formulae for multi-


plication of quaternions.

3.4 Subtraction of finite rotations


Finite rotations possess the property of associativity. This means that the
sequence of rotations 9 1,9 2 ,93 can be performed by composition of the
resultant rotation of 9 1 and 9 2 followed by 9 3 or by composition of 9 1
followed by the resultant rotation of 9 2 and 9 3 •
With this in mind, let us consider the following sequence of rotations:
-9 1,91, 9 2 which is obviously equivalent to 9 2 • This rotation is equivalent
to the sequence of rotations -9 1 and 9. Therefore, replacing in eq. (3.13)
vectors 9,9 1,9 2 by 9 2 , -91, 9, respectively, we arrive at the formula for
subtraction of rotations

92 = 11 (9 - 91 + ~91 X 9) , (3.4.1)
1 + 491.9 2
enabling the second rotation to be determined from given resultant and
first rotation.
Given the resultant and the second rotation the first rotation can be
determined as follows. The sequence -9 2 and -9 1 is equivalent to the
resultant vector -9. For this reason, using eq. (1) we obtain

or

(3.4.2)

3.5 Commutative finite rotations


The above conclusion that the rotations are not commutative was made
under assumption that the directions of the rotation axes are fixed in the
basis system which is not bound to the body. Abandoning this assump-
tion we arrive at conclusions which considerably simplify the forthcoming
3.5 Commutative finite rotations 123

application of the theory of finite rotations. We prove now the following


theorem [57], that the sequence of finite rotations described by the vectors

() Xl () 2 X2
tan 2
1 = 2el tan 2' = 2e2 (3.5.1)

is equivalent to the sequence

()~ = (}2 = 2e2 tan ~2 , ()~ = 2e~ tan ~l , (3.5.2)

where e~ stands for unit vector into which rotation (}2 carries e.
This vector is determined by eq. (1.11)

(3.5.3)

and then

One can add vector (}2 to the expression in parentheses, then by virtue of
eq. (3.13) we obtain

(3.5.4)

Now we will construct the resultant rotation (}f of the rotation sequence
(2). Equation (4) leads to

(3.5.5)

Then by means of eq. (3.113) we obtain

(}f

Then we add the vector


124 3. Theory of finite rotations of rigid bodies

which is identically equal to zero into parentheses in the latter equation.


Applying eq. (3.13) we can write

(J' = (J + (1 - 1
4(Jl . (J2
1) (1 + 4(J2
1 2) (J2 X [(1 - -41(Jl . (J2) (J-

(1+ ~(J~ ) (Jl - ~ (1 - ~(Jl . (J2) (J2 X (J]. (3.5.6)

Replacing (J in the square brackets by the corresponding expression from


eq. (3.13) we obtain the vector

(Jl + (J2 + ~(J2 X (Jl - (1 + ~(J~ ) (Jl - ~(J2 X (Jl - ~(J2 X ((J2 x (Jl)

= (J2 (1 - ~(Jl . (J2) ,

which is collinear to (J2. For this reason, substitution into eq. (6) yields
(J = (J', (3.5.7)
which proves equivalence of rotation sequences (1) and (2).

3.6 Finite rotation and in terms of Euler's angles


The final position of the trihedron of unit vectors i~, fixed in the body, is
obtained from its initial position described by the trihedron fixed is space
is by a sequence of three rotations

(J3 = 2i~ tan ~ (3.6.1)


through angles 'lj;, {), <p about axis Oz, the nodal line n and axis Oz', respec-
tively. An expression for the resultant rotation can be obtained by applying
twice the formula for composition of rotations. However we will avoid this
calculation by using the theorem on commutative rotations. Rotation (J2
brings vector i3 into coincidence with vector i~. Thus, by virtue of the
above theorem, the sequence of rotations (1) is equivalent to the following
sequence

(J'
2 =
'lj;
2"13 t an 2' (J3 = 2i~ tan~. (3.6.2)
But the two latter rotations occur about the same axis. According to eq.
(3.15) they are replaced by one rotation through angle 'lj; + <p, i.e.

(J~ = 2i~ tan 'lj;; <p • (3.6.3)


3.7 Applications of formula for finite rotation 125

Since n . i~ = 0 and i~ x n = n' we obtain, by means of eq. (3.13),

()
2 ( ntan"2
{j., 'I/J+rp
+ 13 tan- , {j rp+'I/J) . (3.6.4)
= 2- +n tan "2 tan-2-

Projecting this equation on axes i~, i;, i~ we have

() = 2[i~tan%(cosrp+sinrptanrp;'I/J)+ (3.6.5)

.,
12 {j (
tan "2 .
- sm rp 2
rp + cos rp tan - + 'I/J)
- + 13., tan -
'I/J 2
+ -
rp]

2 (., {j 'I/J - rp., {j. 'I/J - rp .,. rp + 'I/J)


rp+'I/J 11 tan "2 cos-2- + 12 tan "2 sm-2- + 13 sm - 2 - .
cos--
2

In this expression the coefficients of i~ are projections of the rotation vector


() on axes Ox'y' z' as well as on axes Oxyz. The magnitude of the rotation
vector X is determined byeq. (2.8). This leads to

and

2 x
cos ~
2{j 2rp+'I/J
cos - = cos -cos - - = cos !!.Icos rp + 'I/J I.
2 2 2' 2 2 2

The sign of parameter ).0 can be taken arbitrarily. For this reason we obtain
from eqs. (5), (2.6) and (2.8) that

).0 = cos % cos _rp_;_'l/J_, ).1 = sin % cos _'l/J_;_rp, }


(3.6.6)
, . {j . 'I/J-rp, {j . rp+'I/J
/\2 = sm "2 SIn - 2 - /\3 = cos "2 sm - 2 - '

These are expressions for the Rodrigues-Hamilton's parameters in terms of


Euler's angles.

3.7 Applications of formula for finite rotation


3.7.1 Rotor in Cardan's suspension
It is clear that position of the axis of the rotor in Cardan's suspension
does not depend upon the sequence of gimbal rotations. Let us prove this
by applying the theorem on commutative rotations. Let el, e2, e3 denote
126 3. Theory of finite rotations of rigid bodies

FIGURE 3.4.

the trihedron of axes bound to the platform of the suspension, el being


directed along the rotation axis of the outer gimbal. Also, h , i 2, i3 (i2 = el)
and i~ , i 2, i~ (i2 = i 2) designate the trihedrons of axes bound to the outer
and inner gimbals, respectively. The axis of rotation of the rotor in the
inner gimbal has direction i~ . Initially all trihedrons coincide.
It is necessary to prove the equivalence of the sequences of rotations , i.e.

(3.7.1)

and

(J' (3
2"12 tan 2' (J'2 Ct ,
= 2"11 tan 2 (3.7.2)
1 =

since i~ is the position of vector il after rotation (J~ = (J2.


The resultant rotation of sequence (1) is given by

. t Ct ., t (3 . , .
2 ( 11 (3 )
an 2 + 12 an 2 + 12 X 11 tan 2 tan 2 '
(J Ct
= (3.7.3)

whereas that of sequence (2) is

(3
(J'
=
2 (.,
12 tan 2 + 11., tan 2
Ct
+ 11., .,
X 12
Ct (3 )
tan 2 tan 2 . (3.7.4)

It is necessary to prove whether the equality

.
11 + 12., . t
X 11 an 2(3 = 11., + 13., t an 2(3
holds. The position of the vectors in this equation is shown in Fig. 3.4 (as
seen from the end of i2) . One can see that

h = i~ cos (3 + i~ sin (3, i; x il = i~ sin (3 - i~ cos (3,


3.7 Applications of formula for finite rotation 127

-
s

FIGURE 3.5.

and the problem is reduced to proving the following identities

cos {3 + sin {3 tan ~ = 1, - tan ~ cos {3 + sin {3 = tan ~.

Rodrigues-Hamilton's parameters of rotation () equal

Vo a cos 2'
= cos 2 {3 VI .a
= sm 2 cos {3}
2'
(3.7.5)
a.{3 .a.{3
V2 = cos 2 sm 2' V3 = sm "2 sm "2 .
The same expression can be obtained from eq. (3.18) when we notice that,
due to eq. (2),

Ao = cos a'
{3
Al = 0,
. a
\
/\2
. {3
= sIn 2'
ILo = cos 2' ILl = sm 2 , IL2 = 0,

By means of Table 1 of this chapter and eq. (5) it is easy to reconstruct


matrix (2.6.2).

3.7.2 Rotation of the geocentric system of axes


The definition of this system of axes is given in Sec. 2.15. Let IPo and Ao
be the northern latitude and eastern longitude of the origin of the system
at an initial time instant t = to where IP and A are those at time instant t.
The rotation vector which makes the initial position coincide with the final
position is sought.
128 3. Theory of finite rotations of rigid bodies

We will consider composition of the rotations

~A
O2 = 2etan 2. (3.7.6)

Here e2 is the unit vector normal to the meridian plane and pointing east,
e is the unit vector of the rotation of the earth and ~A = A - Ao + U (t -
to), U being the angular velocity of the earth rotation, see Fig. 3.5. The
first rotation of this sequence corresponds to the move along the meridian
Ao from the point with latitude <Po to the point with latitude <P and the
second rotation along the parallel of latitude <P. By virtue of the theorem
on commutative rotations the sequence (6) can be replaced by

, ~A , , <P - <Po
01 = 2etan 2' O2 = 2e2 tan 2 ' (3.7.7)

where e~ denotes the unit vector perpendicular to the meridian plane A


pointing west. Rotation O2 = O~ carries e2 into e~. Sequence (7) first implies
displacement along the parallel of latitude <Po and then along the meridian
A to the parallel <P. Since the final position of the trihedron does not depend
on the order of our mental constructions, this example readily illustrates
the theorem on commutative rotations.
We have

o = 2 ( e2 tan <P -2 <Po + e tan 2~A + e x e2 tan 2~A tan <P -2 <Po ) '
}

,
o= (~A
2 e tan 2 + e~ tan
<P - <Po
2 + e~ x e tan 2
~A
tan
<P - <Po )
2 .
(3.7.8)

The identity 0 = 0' can be proved by analogy with the previous example.
Since

where e3 denotes the unit vector of the upward vertical and el is directed
north, we obtain

o 2 ( <P + <Po ~A
<P - <Po el cos 2 tan 2 +
cos 2
<P - <Po <P - <Po ~A)
e2 sin 2 + e3 sin 2 tan 2 . (3.7.9)
3.7 Applications of formula for finite rotation 129

The Rodrigues-Hamilton parameters are as follows


<P - <Po ~..\
..\0 = cos 2 cos T'
<P + <Po . ~..\
..\1 = cos 2 sm T'
. <P - <Po ~..\
(3.7.10)
..\2=sm 2 cosT'
. <P + <Po . ~..\
..\3 =sm 2 smT'

3.7.3 Orientation of the axis of a balanced rotor in Cardan's


suspension relative to the geocentric system of axes
The axis of the balanced rotor is known to maintain a given direction in
space. We will use a geocentric coordinate system Oxyz with the origin
o at the start point (<po,..\o). This direction is described by an initial (at
time instant t = to) azimuthal angle ao (angle counted between e1 and
e2) and an initial ascent angle (30 over the horizontal plane. Here ao is
the rotation angle of the outer gimbal of the suspension about the upward
vertical and -(30 is the angle of rotation of the inner gimbal about its axis
in the horizontal plane and initially (ao = 0) pointing west. Required are
formulae for angles a and (3 of the rotor axis relative to the geocentric axes
Ox'y' Zl at the actual place of latitude <P and longitude ..\ at time instant t.
The coordinates of the end of unit vector of the rotor axis are given by

x = cos ao cos (30' y = sin ao cos (30, z = sin (30,


x' = cos cos (3, y' = sin a cos (3, z = sin (3

in coordinate systems Oxyz and OX'yIZI, respectively. Using eq. (2.17) we


obtain
cos a cos (3 = (..\5 + ..\i - -
..\~ ..\~) cos ao cos (30+
2 (..\0..\3 + ..\1..\2) sinao cos (30 + 2 (-..\0..\2 + ..\1..\3) sin (30'
sin a cos (3 = 2 (-..\0..\3 + ..\1..\2) cos ao cos (30+
(3.7.11)
(..\5 + ..\~ - ..\i -
..\~) sin ao cos (30 + 2 (..\0..\1 + ..\2..\3) sin (30'
sin (3 = 2 (..\0..\2 + ..\1..\3) cos ao cos (30+
2 (-..\0..\1 + ..\2..\3) sinao cos (30 + (..\5 +..\~ -..\i - ..\~) sin (30'
where "\0, "\1, ..\2,..\3 are due to eq. (10). These formulae simplify in partic-
ular cases. For example, provided that the rotor axis was initially on the
horizontal plane and pointing east we have ao = - ~ 7r, (3 = 0 and formulae
(11) take the form

cos a cos (3 = - sin~..\ sin <P, sin a cos (3 = - cos ~..\, }


(3.7.12)
sin (3 = sin Ll..\ cos <P.
130 3. Theory of finite rotations of rigid bodies

When the rotor axis was initially on the plane of the meridian (ao = 0)
then calculation with the help of eq. (11) yields

cos a cos (3 = cos <I> cos (<1>0 - (30) + cos ~A sin <I> sin (<1>0 - (30)' }
sin a cos (3 = - sin ~A sin (<1>0 - (30) ,
sin (3 = sin <I> cos (<1>0 - {30) - cos ~A cos <I> sin (<1>0 - (30) .
(3.7.13)

The most simple case is that is which the rotor axis is directed to the
north, then <1>0 = (30 and, by virtue of eq. (13), we obtain

cos a cos (3 = cos <I> , sin a cos (3 = 0, sin (3 = sin <I> ,

which implies a = 0, (3 = <1>, i.e. the rotor axis maintains the direction of
the earth axis rotating about the axis of the inner gimbal through the angle
equal to a change in the latitude. More details can be found in [66].

3.8 Expressions for the angular velocity vector in


terms of finite rotation
The position of a rigid body, that is the position of axes Ox' y' z' fixed in the
body, is given by the vector of finite rotation 8 making axes Oxyz coincident
with axes Ox' y' z'. A position which is infinitesimally close to the actual
position can be obtained in two ways: either by subjecting the body to an
infinitesimal rotation 8' 8 from the actual position or by subjecting the body
to a rotation 8 + 88 from the initial position. Here the infinitesimal vector
88 is an increment in 8 due to the addition of an infinitesimal rotation
such that 8 + 88 is the resultant rotation of two consequent rotations:
finite 8 and infinitesimal 8'8. For this reason, the latter can be found by
the formula of subtraction of rotations (4.1). Neglecting terms of second
order this formula yields

8, 8 = 11 2 ( 88 + -8
1 x 88) . (3.8.1)
1 + 4(} 2

It is easy to obtain an expression for 88 in terms of 8'8 by means of eq.


(4.2)

88 8'8 + ~8'8 x 8 + ~88.8'8

8'8 (1 + ~(}2) - ~8X (8'8 + ~8'8 x 8) . (3.8.2)

The angular velocity vector w is related to the vector of infinitesimal


rotation by equality wdt = 8'8. Making use of eq. (1) yields the equation
3.8 Expressions for the angular velocity vector in terms of finite rotation 131

relating w to the vector of finite rotation

w = 11 2 ("() + "2(}
1 x ()") . (3.8.3)
1 + 48

The time derivative of the rotation vector is expressed in terms of the


angular velocity vector by means of the formula
" 1 1
() = w+-w
2
x (}+-(}() . w
4 . (3.8.4)

Differentiating expression (2.10) for () and taking into account eq. (2.8)
we obtain

where i 1, i 2, i3 are the unit vectors of the fixed axes Oxyz. Due to eq. (3)
the projections of vector w on these axes are

~1 = 2 (AO~l - A1~0 + A2~3 - A3~2)' )


W2 = 2 ( AOA2 - A2AO + A3A1 - A1A3) , (3.8.6)
W3 = 2 ( AO~3 - A3~0 + A1~2 - A2~1 ) .

Projecting now eq. (2) on axes Oxyz yields the relationships

2 [AO~l - A1~ol = ~1 (A~ + Ai) + W2 (AO~3 + ~lA2)2+ W3_ (-AOA2 + A1 A3)


2 AOA2 - A2AO = W1 (-AOA3 + A1A2) + W2 (AO + A2) + W3 (AOA1 + A2A3)
2 AO~3 - A3~0 = W1 (AOA2 + A1A3) + W2 (-AOA1 + A2A3) + W3 (A~ + A~)
(3.8.7)
Using constraint equation (2.7) and its consequence in the form

(3.8.8)
we easily find expressions for the time-derivative of Rodrigues-Hamilton's
parameters in terms of projections of the angular velocity on the fixed axes

2~0 (W1A1 + W2 A2 + W3 A3) , )


= -
2~1 = W1AO + W2A3 - W3A2'
(3.8.9)
2~2 = W2AO + W3A1 - W1A3,
2~3 = W3AO + W1A2 - W2A1.
132 3. Theory of finite rotations of rigid bodies

Given these projections we can find projections on axes Ox'y' z' fixed in
the body by means of Table 1 of direction cosines (Sec. 3.2). The result is

WI = 2 ( AO~1 - AI~O + ~2A3 - ~3A2 ) , )

W2 = 2 ( AOA2 - A2AO + A3AI - AIA3) , (3.8.10)

W3 = 2 ( AO'x3 - A3'xO + 'xIA2 - 'x2AI) .

The sign of the second group of terms has changed. This could be pre-
dicted as coordinate systems Ox'y'z' and Oxyz exchange places under ro-
tation -(). This means that the signs of AI, A2, A3 and w changes whilst the
sign of AO is unchanged. Expressions for the time-derivatives of Rodrigues-
Hamilton's parameters in terms of Ws take the form

2~0 = - + W2A2 + W3A3),


(WIAI }
2AI = WIAO + W3A2 - W2A3, (3.8.11)
2'x2 = W2AO + WIA3 - W3Al,
2'x3 = W3AO + W2AI - WIA2.

As an example ofthe application offormula (3) we determine the angular


velocity of the geocentric system of axes whose rotation vector () is given
by eq. (7). We have

1 + -()
1 2
= cos2 II> - 11>0 cos2 -~A
4 2 2
Differentiating with respect to time and taking into account that vectors
e and e2 do not move we obtain
<i> ~,x
iJ = II> - 11>0 e2 + ~A e +
cos 2 cos2 -
2 2

( ~,x
--~'A~ tan
cos 2 -
II> - 11>0
2 +
cos 2
<i>
II> - 11>0 tan""2
---
~A) e*,
2 2
where e* = e x e2 is a unit vector directed outward along the radius of the
parallel circle, e2, e*, e forming a right-handed trihedron of unit vectors.
Inserting this expression into eq. (3) we have
(3.8.12)
Projecting this equality on the axes of the geocentric system we obtain

w U + el (,x cos 11>0 - <i> sin ~A sin 11>0) +


e 2<i> cos ~A + e3 (,x sin 11>0 + <i> sin ~A cos 11>0)' (3.8.13)
3.9 Cayley-Klein's parameters 133

~J
N

FIGURE 3.6.

where el,e2,e3 denote the unit vectors of the geocentric system (that is
the northern and western directions and upward vertical) of the starting
position at the instant of the start, and U =U e designates the angular
velocity of the earth.
Expressions (12) and (13) can be also represented in the form

w = e~>' + e~ <i> = U + >. (e~ cos <I> + e~ sin <I» + e~ <i>,


where e~ , e;, e~ are the trihedron of the geocentric system of the actual
place at time instant t.

3.9 Cayley-Klein's parameters


These parameters, defining position of a rigid body, present complex-valued
combinations of Rodrigues-Hamilton's parameters. Using these parameters
a rational transformation in the complex plane is found for any rotation
and the problem of rotation composition reduces to a sequence of these
transformations.
In what follows , we use two representations of a complex number: first
by a point z = x + iy in the complex plane and second by means of a
stereographic projection of a point on a unit sphere (Riemann's sphere)
onto the plane.
Let plane z denote the equatorial plane of the sphere Eo of unit radius.
Points in the plane are projected stereographically from the pole N on the
sphere Eo, that is points N, M and M' lie along a straight line, see Fig.
3.6. The fixed axes O~1~2~3 are taken so that the origin lies at the centre
of the sphere, axes ~l and ~2 coincide with axes Ox and Oy, respectively,
and axis O~3 intersects the sphere at the pole N. Then, the fact that the
three points N(O,O , 1) , M(~1 ' ~2 ' ~3) and M(x,y,O) lie along the straight
134 3. Theory of finite rotations of rigid bodies

line N M M' is expressed as follows

x Y 1
I.e. Z
. = ~l + i~2
= X + zy (3.9.1)
~l ~2 1- ~3' 1- ~3 '

the quantities ~l' ~2 and ~3 being related by the equation for the unit sphere

d + ~~ + .;~ = 1. (3.9.2)

Stereographic projection transforms the circles lying in the plane P and


containing the sphere centre into great circles on the sphere. Indeed, by
virtue of eqs. (1) and (2), the following expression

is transformed into a circle equation

B) 2
( X - 2A ( C ) 2 B2 + C2
+ y - 2A = 1 + 4A2

in the plane

J+
passing through the origin 0 of the coordinate system.
2
This circle contains the sphere centre since the radius R = 1 B 24: 2C
of the circle is greater than the distance between the centres of the circle

and the sphere which is equal to JB +C2 2


4A2' It is easy to show (however
we omit the proof) that stereographic projection of any circle in the plane
P is a circle on the sphere. In particular, straight lines in plane P, passing
through the coordinate origin, are transformed into great circles passing
through the poles Nand S. The unit circle Izl = 1 corresponds to the
equator, points within the unit circle (i.e. Izi < 1) correspond to points on
the lower hemisphere (~3 < 0) and points outside the unit circle (Izl > 1)
correspond to points on the upper hemisphere (~3 > 0). The origin z = 0
and the infinite point of plane P correspond to the poles Sand N, respec-
tively.
We turn now to the case of a rigid body with a fixed point 0 and imagine
a sphere ~ circumscribed about 0 and rigidly bounded to the body and
axes OXIX2X3' During the body rotation the sphere ~ slides over the fixed
sphere ~o. The rotation changes the great circle r of the sphere ~ into
circle r ' on ~o. Let "( and "(' be the circle in plane P corresponding to the
3.9 Cayley-Klein's parameters 135

great circles rand r'. This change of , to " corresponds to a rotation


of the body which is equivalent to a rotation of the sphere ~. Under the
above change the complex-valued coordinates of the points of circles, and
" are related to each other by means of a rational transformation
, az + (3
z =---. (3.9.3)
,z + t)
The coefficients a, (3", 8 of this transformation are determined up to a
constant factor since relationship (3) does not change when each term in
the numerator and the denominator is multiplied by a number. That is why
the choice of the coefficients is subject to a normalisation condition taken
in the form

at) - {3, = 1. (3.9.4)


The quantities a, {3", 8 are referred to as Cayley-Klein's parameters, [50].
The relation between Cayley-Klein's parameters and Rodrigues-Hamilton
parameters is required. To find it we observe that a rational transformation
bringing three points ZI, Z2, Z3 of circle , to three points z~, zb, z~ of circle
" can be written in the form
z - Z2 Z3 - ZI Z' - zb z~ - z~
(3.9.5)
Z - ZI Z3 - Z2 z, - z~ z~ - zb .
The rotation vector () intersects sphere ~o at point P, whose coordinates
are proportional to the Rodrigues parameters },1, },2,},3 and at a diametri-
cally opposite point Q. The proportionality coefficient should be chosen so
as to satisfy the sphere equation (2). For this reason, due to eq. (2.7), we
obtain expressions for the coordinates of points P and Q in the form

In view of eq. (1) the corresponding points P' and Q' in plane P have the
coordinates

Since points P and Q does not move under the body rotation, we can set

in the coordinate transformation (5). Point ZI = 00 corresponded to the


pole N before the rotation. The coordinates ~1' ~2' ~3 of this point after
rotation are
136 3. Theory of finite rotations of rigid bodies

which follows from eq. (2.18) by taking Xl = X2 = O. Taking into account


eq. (2.7) we obtain, by means of eq. (1), that
, A3 - iAo
zl = .
Al - iA2
Substituting

)1- A6 + A3
Zl = 00,

into eq. (5) we have

Z'
Z(A3 - iAo) + Al + iA2
Z (AI - iA2) - (A3 + iAo) ,
and thus

The proportionality coefficient C is determined by eq. (4). The final


expressions for Cayley-Klein parameters in terms of Rodrigues-Hamilton
parameters are given by

One can see that a and 8, as well as (3 and '"Yare complex conjugates.
Moreover, lal 2 + 1(31 2 = 1112 + 181 2 = 1 due to eq. (4).
Due to eq. (6.6) expressions for Cayley-Klein parameters in terms of
Euler's angles take the form

a = {) (.7fJ+tp) , (3 = zsm2"ex
cos 2" exp z-2- .. {) p (.7fJ-tp)
z-2- , }
(3.9.7)
8= {) (.7fJ+tp) .. {) (.7fJ-tp)
cos 2" exp -z-2- , '"Y = z sm 2" exp -z-2- .

The transformation of coordinates (Xl, X2, X3) into (~l' ~2' ~3)' expressed
byeq. (3), can be written in the form

(~1'~2'~3) = (~ ~) (Xl,X2,X3). (3.9.8)

However it follows from eq. (3) that


-8z' + (3
z=---- (3.9.9)
,,(z' - a
3.9 Cayley-Klein's parameters 137

Thus, the inverse transformation, i.e. transformation of (~l' ~ 2, ~ 3) into


(XI,X2,X3) is as follows

(3.9.10)

An explicit expression for transformation (8) due to eq. (1) is

a (Xl+ iX2) + fJ (1 - X3) (3.9.11)


,(Xl +iX2) +8(I-x3)·
The complex conjugate is then
8 (Xl - iX2) - ,(1 - X3)
(3.9.12)
-fJ (Xl - iX2) + a (1 - X3)"
Multiplying eqs. (11) and (12) we obtain by virtue of eqs. (2) and (4) that

1 + [(a8 + fJ,) X3 - (a,- fJ8) Xl - (a, + fJ8) iX2J


1 - [(a8 + fJ,) X3 - (a, - fJ8) Xl - (a, + fJ8) iX2J
and conclude that the expression in the square brackets is ~3. Using this
expression for ~3 we obtain ~l + i~2 and ~l - i~2 and arrive at the trans-
formation formulae

~l + i~2 + iX2) - fJ2 (Xl - iX2) - 2afJx3,


= a 2 (Xl }
~l - i~2 = _,2 (Xl + iX2) + 82 (Xl - iX2) + 2,8x3, (3.9.13)
~3 = -a, (Xl + iX2) + fJ8 (Xl - iX2) + (a8 + fJ,) X3·

The inverse transformation is obtained by using eq. (10)

Xl + iX2 = + i~2) - fJ2 (~l - i~2) + 28fJ~3'


82 (~l }
Xl - iX2 = _,2 (~l + i~2) + a 2 (~l - i~2) - 2,a~3' (3.9.14)
X3 = ,8 (~l + i~2) - fJa (~l - i~2) + (a8 + fJ,) ~3·
Dealing with equations (13) and (14) for the linear transformation there
is no need to think that the variables are normalised in accordance with
eq. (2).
When one uses Cayley-Klein's parameters the problem of rotation com-
position reduces to performing a sequence of rational transformations, namely
the transformation

(3.9.15)

corresponds to a first rotation (h, the next transformation

z" a2 z ' + fJ2


'2z' + 82
138 3. Theory of finite rotations of rigid bodies

corresponds to a second rotation (}2.


The transformation z into z" is given by

Ct12Z + f3 12 (3.9.16)
1'12 z + 812 .

and describes the resultant rotation ().


Thus, Cayley-Klein's parameters of the resultant rotation are expressed
in terms of rotations (}1 and (h as follows

Ct12 = CtlCt2 + 1'1f32, f312 = Ct2f31 + 81f32, }


(3.9.17)
1'12 = 1'2 Ct l + 1'1 82, 812 = 1'2f3 1 + 81 82 .

It is easy to prove that these formulae are in agreement with eq. (3.18).

3.10 Angular velocity in terms of Cayley-Klein's


parameters
To begin with, we denote Cayley-Klein's parameters at time instants t and
t + dt as

Ct, f3, 1', 8·, and Ct + adt, f3 + iJdt, l' + /ydt, 8 + 8dt,

respectively. In eq. (9.17) they can be understood as parameters Ctl, f31, 1'1' 8 1
of the first rotation and parameters Ct12, f312' 1'12' 812 of the resultant rota-
tion. In the case under consideration the second rotation with parameters
Ct2, /32, 1'2' 82 is an infinitesimal rotation 8'(} =wdt whose projections on axes
O~I~2~3 are equal to

- d
WI t = 2AI
Ao'

Then with accuracy up the to second order we can adopt

L
AO = 1, Al = 2WIdt, (3.10.1)

and, due to eq. (9.6),

(3.10.2)
3.11 Determination of a rigid body position from angular velocity 139

(3.10.3)

etc. Then we find

a = ~W3a + ~ (WI + iW2) 'Y, iJ = ~W3,8 + ~ (WI + iW2) 8, }

1 = -~W3'Y + ~ (WI - iW2) a, 8 = -~W38 + ~ (WI - iW2) 'Y.


(3.10.4)

Replacing here the projections Ws of the angular velocity vector on axes


0';1';2';3 by their expressions in terms of projections Ws on axes fixed in
the body with the help of formulae (9.14) we obtain

(3.10.5)

It is easy to construct the inverse relationships

(3.10.6)

and

WI + iW2 = 2i (,88 - iJ8) , }


WI - iW2 = 2i ("fa -1a), (3.10.7)
W3 = 2i ( a8 - iJ'Y) = 2i (,81 - a8) .

3.11 Determination of a rigid body position from


angular velocity
It is assumed that projections WI, W2, W3 of the angular velocity vector
of the body rotating about a fixed point on axes fixed in the body are
prescribed functions of time. Time-dependence of parameters describing
140 3. Theory of finite rotations of rigid bodies

the body position is sought. Thus we speak about integration of the system
of differential equations providing us with expressions for derivatives of the
above parameters in terms of Wl,W2,W3.
A first form of these differential equations is given by expressions (2.10.1)
for derivatives of Euler's angles in terms of projections Ws of the angular
velocity vector

fJ=WICOScp-W2sincp, }
. 1
'ljJ = --:--::a (WI sin cp + W2 cos cp) , (3.11.1)
Slnu
(p = W3 - (WI sin cp + W2 cos cp) cot fJ.

Certain of the particular cases admit integration of this system. How-


ever nonlinearity and the absence of symmetry complicate the solution and
prevent general results being obtained.
A symmetrically constructed linear system of differential equations can
be obtained by introducing into consideration a unit vector e keeping a
constant direction in space. Denoting its projections 'Y 1, 'Y 2, 'Y 3 on axes fixed
in the body, i.e. the direction cosines of angles between e and axes Ox' y' z',
we have

As the velocity of the end of vector e is equal to zero, we can write, due to
eq. (2.13.3),
3

e = L h si~ + 'Y sW x i~) = O.


s=1

It follows that the coefficients of i~ vanish which yields a system of three


differential equations

1'1 + W2'Y3 - W3'Y2 = 0, 1'2 + W3'Y1 - WI 'Y3 = 0, 1'3 + WI 'Y2 - W2'Y1 = O.


(3.11.2)

An obvious consequence of these ~quations is the following relationship

and its integral

(3.11.3)

Also the initial values 'Y~, 'Yg, 'Y~ should satisfy this condition. Thus the
problem is reduced to integration of the system of three differential equa-
tions (2) having the first integral (3).
3.11 Determination of a rigid body position from angular velocity 141

Let us assume that the solution depending upon two integration constant
is found and we take e = i1 which implies that vector e is directed along
axis Ox. Then we can determine the integration constants by means of
the initial conditions 'Y~ = a~l' 'Yg = ag1' 'Y~ = a~l satisfying condition (3)
where a~l denote prescribed direction cosines of the angles between axes
Ox' y' z' and axis Ox at the initial time instant. We obtain all, a21, a31
which yields the first column in Table 1 of directions cosines of Chapter 2.
By analogy we can find the second and the third columns of this table.
Another way of solving the problem is to find Rodrigues-Hamilton's pa-
rameters. In this case we need to integrate a system of four differential
equations (8.11) having the first integral (2.7). If we introduce complex-
valued combinations (9.6) of Rodrigues-Hamilton's parameters, i.e. Cayley-
Klein's parameters, then system (8.11) takes form of (10.5) and the fin;t
integral (2.7) the form of (9.4). Using Cayley-Klein's parameters leads to a
simpler statement of the problem. Indeed, system (10.4) is then split into
two systems of linear equations of the first order and of completely similar
structure. Both have the form
. iW3 1 .
P = - - P - - (W2 - zwd u. (3.11.4)
2 2
Let us assume that two linear independent solutions of this system, form-
ing the unit matrix

(3.11.5)

are known and satisfy the initial conditions. The solution of system (10.5)
subject to the initial conditions

t = 0 a = ao, (3 = (30, 'Y = 'Yo, 8 = 80, (3.11.6)


which satisfy the following relationship

a 0 80 - (3o'Yo = 1 (3.11.7)

can be cast in the form

a = aOU1 + (3ou2, (3 = aOP1 + (30P2, 'Y = 'YO U 1 + 80X2' 8 = 'YOP1 + 80P2'


(3.11.8)
All we need is to prove that this solution satisfies condition (9.4). Recall
that the system of the differential equations

x = Pll (t) x + P12 (t) y, if = P21 (t) X + P22 (t) Y


has the Wronskian

D(t) ~ x,y, -x,y, ~ D (0) exp ( - / (PH + p-"Jdt)


142 3. Theory of finite rotations of rigid bodies

But for system (4) Pu + P22 = 0 and for the assumed solutions D (0) = 1.
Thus

and, in view of (8),

which completes the proof.


Actually, it is sufficient to find only one system of particular solutions
Xl, Pl' The second system is obtained from the equalities

where a bar denotes the complex conjugate. Indeed, Xl and PI are solutions
of the system of differential equations obtained from (4) by replacing i with
-i in the right-hand side. In addition to this, Xl = 1 and PI = O. These are
the functions X2 and P2 which satisfy these equations and the boundary
conditions.

3.12 Darboux's equation


In the classical treatise by Darboux [23] on the theory of surfaces, the prob-
lem of determining position of a body for given angular velocity is reduced
to finding a particular solution of a Rikatti-type equation. Derivation of this
equation is based upon consideration of stereographic projection of a plane
on a unit sphere ~o, the problem being discussed in Sec. 3.9. Let ~1'~3'~3
designate the coordinates of a point on the sphere. Its coordinates relative
to axes OXIX2X3 are given by transformations (9.10) and (9.9). Differenti-
ating the latter equation with respect to time and taking into account that
Z' = 0 we obtain

Z= 1
hz' - 0')
2 [(-bz' +~) hz' - 0') - (-l5z' +(3) (i'z' - a)] .
Applying formulae (10.6) we have

Z= 1 [WI + iW2 + 2Z'W3 - z'2 (WI - iW2)] . (3.12.1)


2i hz' - 0')2

Since, due to eq. (9.3),


1 Z'
- ("(Z + ,6) , -----=2 = (O'z + (3) h z + 15) ,
hz' - 0')
(O'z + ,6)2 ,
3.12 Darboux's equation 143

we use formulae (9.14) for transformation of vector w

WI + iW2 = 82 (WI + iW2) - /32 (WI - iW2) + 2/38w3, }


WI - iW2 = -,,? (WI + iW2) + a (WI - iW2) - 2')'aW3,
2 (3.12.2)
W3 = ')'8 (WI + iW2) - /3a (WI - iW2) + (a8 + /3')') W3,

and deduce the Darboux-Rikatti equation

. W2 - iWI. W2 + iWI 2
Z = 2 - ZZW3 + 2 z. (3.12.3)

The form of this equation can be simplified. Using the substitution

(3.12.4)

U
the linear term on the right-hand side can be eliminated

i; ~ ~ (w, - iw,) exp w3dt) + ~ (w, + iw,) " exp ( !


-i w3dt) .
(3.12.5)

Next assume

!
so that

!1~ vwl+wl, q~expi [",g(W,+iwd - w3 dt],


1
q= -.
q
(3.12.6)

Now introducing a new independent variable

(3.12.7)

which increases monotonically as t increases, we cast the differential equa-


tion (5) in the form

(3.12.8)
144 3. Theory of finite rotations of rigid bodies

where a prime denotes differentiation with respect to time.


We demonstrate now that the problem of determining the position of
a rigid body rotating about a fixed point is completely solvable provided
that one particular solution of this equation is found. Indeed, if ( = ~ is a
particular solution of this equation then the second particular solution is
-1/~. Then it follows from a well-known property of the Rikatti equation
that the function
(-~ (-~­
rJ = - - 1 = (~+ 1 ~ (3.12.9)
(+ -
~
satisfies a linear differential equation of the first order. This equation is
easy to construct. Indeed,

rJ' -('--( - (' + (~)'


rJ (-~ (+~
~
However from the differential equations which is the complex conjugate of
eq. (8) we have

-,
~ = q+
1-2
-~ ,
q t ~
q ~ - ~ q'

and the previous relationship yields

t J(q~ -~)
T _
~
rJ'
- - =
rJ ~
=q~ --,
q
i.e. ~ = Cexp dT = Cexp (2i8(T)) ,
o
where

(3.12.10)

is a real-valued function since it is a difference of two complex conjugated


values divided by 2i. By means of eqs. (4) and (9) we find

z = exp ( -~
.
I
t
W3
d
t)
~+Cexp(2i8(T))
1- C~ exp (2i8(T)) .
Multiplying the numerator and the denominator by the factor

~exp (-i letT) - ~! wed'])


3.12 Darboux's equation 145

and introducing the notation

1 . t;
a - ~ exp (-z8 1 ), b = ~ exp (-i8 2 ),

c -I exp(i8 2), d= - ~ exp(i8 1 ), (3.12.11)

where

J
t t

~=Jl+t;~, 81=8(T)-~JW3dt, 8 1 = 8(T) +~ W3 dt ,


o o
(3.12.12)

we arrive at the expression for the general solution of the Darboux differ-
ential equation in the form of a rational transformation

-dC+b
Z=--- (3.12.13)
cC-a
where the factors -d, b, c, -a are normalised due to eq. (9.4). It is necessary
to find such a value of C that eq. (9.9) and initial conditions (11.6) are
satisfied. To this end we set

C = mz' +n
pz'+q
and require that m, n, p, q meet the conditions

md - pb = 8, -nd + qb = /3, mdo - pbo = 80, -ndo + qbo = /30'


mc - pa = "I, -nc + qa = a, mea - pao = "10, -neo + qao = ao·

Eliminating for example nand q from the equations containing a, ao, /30
On the right hand side, we arrive at the relationship which determines a

a c -a
ao Co -an = o.
/3 0 do -bo

Thus we obtain expressions for Cayley-Klein's parameters in the form

a = ao (ado - cbo) + f3 0 (cao - aeo), }


/3 = ao (bdo - dbo) + /30 (dao - ben) ,
(3.12.14)
"I = "10 (ado - cbo) + 80 (cao - aea),
8 = "10 (bd o - dbo) + 80 (dao - bco) .

The expressions in the parentheses are calculated by meanS of eqs. (11)


and (12). On the other hand, comparing them with formulae (11.8) allows
us to equate them to the corresponding solutions of system (11.4) subject
146 3. Theory of finite rotations of rigid bodies

to initial conditions (11.5). This leads to equations relating the solution


of the above mentioned system to the assumed particular solution of the
Darboux equation

ado - cbo = Xl = ~~o [exp (-i8 l ) + e~o exp (i8 2 )] ,


bdo - db o = PI = ~~o [~o exp (i8 l ) - ~ exp (-i8 2 )],
(3.12.15)
cao - aeo = X2 = 1 [-
~~o ~ exp (i8 2 ) -
-
~o exp (-i8 l )
] ,

dao - beo = P2 = ~~o [exp (i8 l ) + ~eo exp (-i82 )] ,


where subscript 0 denotes values relating to the initial values of the partic-
ular solution ~ and ~o = Jl
+ ~o~o·
Thus, the problem of determining the position of a rigid body rotat-
ing about a fixed point is reduced to the quadrature, provided that one
particular solution of the Rikatti-type equation is known.

3.13 An example. The position of a self-excited


rigid body
We consider a body rotating round a fixed point O. The body is assumed
to be subject to time-dependent moments but not position-dependent mo-
ments. For example, such moments can be achieved by the forces of a
reaction jet. A symmetry of rotation about axis Oz' is assumed, that is
moments of inertia about axes Ox' and Oy' coincide (A = B). The mo-
ment of inertia about axis Oz' is denoted by C.
Equations of rotation of a body about a fixed point, known as Euler's
equations, are easy to integrate. We have

AWl+(C-A)W2 W3=ml' }
AW2 - (C - A) W3Wl = m2, (3.13.1)
CW3 = m3·
The latter equation yields

J
t

W3 = wg + ~ m3 dt , (3.13.2)
o
whereas the first two equations can be combined when we multiply the
second equation by i and add the first. The result is a linear equation of
the first order
. .. C-A. ( .) 1( .)
W2 + ZWI + -A-ZW3 W2 + ZWI = A m2 + zml ,
3.13 An example. The position of a self-excited rigid body 147

whose first integral is given by

[Wl+ Uv 1+ ~l (m,+im,)exp (iC~A I


W3 dt ) &]
exp (_iC ~ A W3&) &. I (3.13.3)

A simple case is that in which there are no excitation moments. This


is the case (for A = B) of regular precession, and the time-dependence of
values defining the position of the body is easy to obtain by applying the
approach of the previous Section. Indeed, for ml = m2 = m3 = 0 we have

W3 = wg, W2 + iWI = (w~ + iw~) exp ( _i C ~ Awgt)

wOexp ( -i [C ~ A wgt + 2c]) , (3.13.4)

where wO and - 2c denote the absolute value and the argument of wg + iw~.
By means of formulae (12.6) and (12.7) we obtain
2T
t= 0 ' q=exp(-2i[TcotA+c]), (3.13.5)
w
where A is given by

Cwg
cot A = Aw o . (3.13.6)

The Rikatti equation (12.8) takes the form

(' = exp (2i [Tcot A + c]) + exp (-2i [TcotA + c]) (2, (3.13.7)

and its particular solution is sought in the form

~ = Aiexp (2i [Tcot A + c]),


where A is a constant which should be determined from the quadratic
equation

A2 - 2A cot A - 1 = O. (3.13.8)

A A.
The roots are Al = cot 2 and Al = - tan 2· Takmg the first root we
obtain the following particular solution

~ = icot ~ exp (2i [Tcot A + c]), (3.13.9)


148 3. Theory of finite rotations of rigid bodies

and calculation with the help of formulae (12.10)-(12.12) yields

8 ( T) = Tcot "2'>. 8 I = T(cot


> "2. c
- A cot>. ) , 8 2 = T (cot ~ + ~ cot >.) .
Then by virtue of eq. (12.15) we find

Xl = ~ exp (iT ~ cot>.) (exp (-iT cot ~) +


sin 2

cot 2 ~ exp (iT tan ~) ) = P2

p, ~ ~ ,in A ex: (-iT [~ oot A - 2£]) (ex (ircot D-


p

exp (-iT tan "2) ) = -):(2

(3.13.10)
In order to simplify the final expression we adopt the zero initial values of
Euler's angles, then by means of eq. (9.7) we obtain
ao = 1, f3 0 = 0, 1'0 = 0, 80 = 1
and moreover
{}
a = cos "2 exp (.1/J+<P)
t -2- = Xl,
.. {} exp t --
(3 = tsm"2
2
(.1/J-<P) = Pl'
(3.13.11)
What remains is to separate the real and the imaginary part of these ex-
pressions.
When a motion differs from regular precession the problem becomes
much more difficult. Consider, for instance, the case of ml = m2 = 0,
but m3 -=I=- O. Keeping the above notation for the independent variable
T and constants w O , E, cot>. and introducing the relative angular velocity
e = W3/wg we rewrite expression (3) in the form
(3.13.12)

Instead of the Rikatti equation we turn to the system of linear equa-


tions (11.4) and introduce new unknown variables k and r by means of the
following relationships

k xexp (i [(1- ~) ootA IOdT+E]) ,


r pexp ( -i [(1 - ~) cot A I l) ,
OdT + £ (3.13.13)
3.13 An example. The position of a self-excited rigid body 149

Then for k and r we obtain the system of differential equations

k' = ike cot A + r, r' = -ire cot A - k, (3.13.14)

where a prime implies differentiation with respect to T.


e
The case = 1 corresponds to the above case of the regular precession.
Obviously, the solution of the obtained system of linear differential equa-
tions leads to relationships (10) under the initial condition x (0) = 1 and
p (0) = O.
The system (14) of two differential equations of first order can be reduced
to a single equation of second order

k" + k (1 + e 2 cot 2 A - ie' cot A) = O. (3.13.15)

In particular, when the moment of forces about axis Oz' is not time-
dependent, variable e depends linearly on time

(3.13.16)

and the problem reduces to the Weber differential equation, [94].


4
Basic dynamic quantities

4.1 Kinetic energy of a system


The kinetic energy of a system of particles is equal to half the sum of the
masses multiplied by the velocities squared

1 N 1 N
~m'v~
T= -2 L...J t t
~m'v'
= -2 L...J t t
·v·t· (4.1.1)
i=l i=l

We obtain an expression for the kinetic energy by replacing the gener-


alised velocities Vi with the expression given in eq. (1.3.3), i.e.

The kinetic energy is then as follows

T =
152 4. Basic dynamic quantities

We now introduce the notation that

Adopting notation (1.2.1) and assuming that

etc. we can also write

To
1
="2 Lmi
3N (8~i)2
at
"=1
(4.1.3)

These values depends upon the generalised coordinates and time t, how-
ever there may exist cases when variable t, being incorporated explicitly in
expressions (1.2.9) for Cartesian coordinates in terms of generalised coor-
dinates, does not appear in Ask, Bs and To.
The kinetic energy can be represented as the sum of three terms

(4.1.4)

The first term, i.e. T 2 , contains the components which are quadratic with
respect to the generalised velocities

(4.1.5)

The second term is linear with respect to the generalised velocities


n
T1 = LBsqs. (4.1.6)
s=l

The last term in (4), i.e. To, does not depend on the generalised velocities.
In the case of stationary constraints, the generalised coordinates are taken
so that t is absent in expressions for the Cartesian coordinates (or positions
vectors ri) in terms of the generalised coordinates. Then it follows from eq.
(2) that

Bs = 0, To = 0, (4.1.7)
4.1 Kinetic energy of a system 153

that is, the kinetic energy of the system subjected to stationary constraints
is a quadratic form of the generalised velocities

(4.1.8)

According to definition (1), the kinetic energy is a non-negative value


which is equal to zero only if the velocities Vi of all points of the system
vanish. But, in the case of stationary constraints, this implies that all gen-
eralised velocities are identically equal to zero. This follows from eq. (1.3.4)
since the deficiency of matrix (1.2.12) is equal to zero. For this reason the
quadratic form (8) of the generalised coordinates is positive definite and
its coefficients satisfy Sylvester's criterion expressing positiveness of the
determinant and all the principal minor determinants corresponding to the
matrix of the coefficients

(4.1.9)

for any values of the generalised coordinates in the domain of definition,


see (A.3.25).
The Sylvester inequalities are as follows

All A 1n
~n = IAI = > 0,
AnI Ann
All A 1,n-l
~n-l = > 0, (4.1.lO)
A n- 1,1 A n- 1,n-l

A121
A22 > 0, ~1 = All > 0.

The first inequality expresses also that matrix A is non-singular, the sec-
ond inequality says that the matrix of coefficients of the quadratic form,
corresponding to the kinetic energy of the system subjected to the con-
straint qn = canst, is non-singular and so on.
Under non-stationary constraints the expression T 2 , eq. (5), is positive
definite form of the generalised velocities. This follows from the fact that
T2 is the kinetic energy of the virtual velocities v~

T2 = ~ ' " m·v,2 = ~ ' "


N
2L..- 2 2 2L..-
N
m.lv. _ariat 12
22
(4.1.11)
i=1 i=1
154 4. Basic dynamic quantities

In what follows we use the Euler theorem on homogeneous functions. A


function cP (Xl, X2,··· ,xn ) of n variables satisfying the condition

is referred to as homogeneous. The theorem expresses the following equality

(4.1.12)

Now we have due to eq. (4)

and by virtue of (12) we have

(4.1.13)

In the case of stationary constraints we obtain the identity

uT
LiJsa:-qs
n
= 2T (4.1.14)
s=l

which is used many times in what follows. We proceed now to construct


an expression for the kinetic energy in terms of the quasi-velocities. We
consider next the case of stationary constraints and consider the quasi-
velocities as being given by homogeneous forms of the generalised velocities.
Then, using the expressions for the velocities

(4.1.15)

we obtain

(4.1.16)

where similar to eq. (2)

N
" " . uri . uri
A *sk -_ ~mt!:\ (4.1.17)
!:\.
i=l uK s uKk
4.2 Associate expression for the kinetic energy 155

If the constraints are non-stationary and the quasi-velocities are defined by


relationships (1.5.22) the velocity vector is

(4.1.18)

Here we obtain
1 n n n
T ="2 LLA;kWsWk + LB;ws +T~, (4.1.19)
s=lk=l s=l
where

(4.1.20)

Expressions for the kinetic energy in terms of the quasi-velocities seem


to have a very complex structure. However in many cases they have much
more simple expressions than those in terms of the generalised velocities.
In what follows we will experience evidence that confirms this.
In the case of stationary constraints the terms with
ar
at t vanish. When the
constraints are non-stationary and the quasi-velocities are introduced by
homogeneous expressions (1.5.1) the terms with coefficients bs ,n+1 vanish.

4.2 Associate expression for the kinetic energy


The quantities known as generalised momenta are of major importance
in analytical mechanics of holonomic systems. The generalised momentum
corresponding to the coordinate qs is denoted as Ps. By definition, it is
equal to the derivative of the kinetic energy with respect to the generalised
velocity ris

aT (4.2.1)
Ps=aris (s=l, ... ,n),
156 4. Basic dynamic quantities

see also (10.2.1). In the case of stationary constraints the generalised mo-
menta are homogeneous linear functions of the generalised velocities
n
Ps = LAskqk, (4.2.2)
k=1
and in the case of non-stationary constraints
n
Ps = L Askqk + Bs· (4.2.3)
k=1
These equations are solvable for the generalised velocities since, as pointed
out above, matrix A is non-singular and thus the inverse matrix A-I exists.
By virtue of (A.2.32) the kinetic energy has the following matrix form

T -- ~2 q., A q. + q., B + T.0, (4.2.4)

where q and q' are a column-matrix and a row-matrix of the generalised


velocities, respectively, and B denotes a column-matrix of coefficients Bs.
Entering a column-matrix p of generalised momenta we obtain instead of
eq. (3)

p=Aq+B, (4.2.5)

which enables an equation for the generalised velocities in terms of the


momenta

(4.2.6)

For stationary constraints we have

p=Aq, (4.2.7)

which leads to the following result

q., =p, (A-I)' ='A-


p, 1 (4.2.8)

as matrices A and A -1 are symmetric. The expression for the kinetic energy
in terms of momenta, denoted as T', takes the form

T' = 1 'A-1AA-1 P
-p 1 'A-1 p.
= -p (4.2.9)
2 2
This expression is referred to as the associate expression for the kinetic
energy.
4.2 Associate expression for the kinetic energy 157

For example, for a system with two degrees of freedom

-A12
All
II ' (4.2.10)

and the associate expression for the kinetic energy is given by

T' - ~ 1
- 2 AllA22 - Ai2
(A22pi - 2A 12 PIP2 + AllP~) . (4.2.11)

We notice also the bilinear representation for the kinetic energy

T 1 ~. I., 1,.
= "2 ~ qsPs = "2 q P = "2P q. (4.2.12)
s=1
In the case of non-stationary constraints one is often interested in the
following value
n
y= LPsrjs - T = 2T2 + Tl - T = T2 - To (4.2.13)
8=1

in terms of the momenta but not in kinetic energy.


The relationship (1.13) is used here. Using the matrix form and eqs. (4)
and (6) it is easy to obtain the associate expression for Y

Y' rj'p-T
(p' - B') A- 1p -"21 (p' - B') A-I (p - B) - (p' - B') A-I B - To

or

Y' = ~2 (p' - B') A-I (p - B) - To . (4.2.14)

In order to obtain a matrix form for the kinetic energy in terms of the
quasi-coordinates, we rewrite eq. (1.5.1) and its inverse in the form

(4.2.15)

Inserting the latter into eq. (4) we have


1 1
T = "2w'b' Abw + w'b' B + To = "2w' A*w + w' B* + To, (4.2.16)

in which the n x n matrix A* and the n x 1 column-matrix B* are given


by the relationships

A* = b'Ab, B* = b'B. (4.2.17)

The first equality in this equation implies another form of eq. (1.17) whilst
the second is eq. (1.20) for bs,n+l'
158 4. Basic dynamic quantities

4.3 Tensor of inertia


While considering the motion of the material system it is important to
introduce the quantities and concepts which characterise the distribution
of the mass of the particles within the system.
The first concept is the centre of inertia which is the geometrical point
whose position vector rc is determined by the equation
1 N
rcM Lmiri.
= (4.3.1)
i=1
Here the mass of the system which is the sum of particle masses
N

M=Lmi (4.3.2)
i=1
is introduced. When the system is moving the position of the centre of mass
changes not only with respect to inertial axes Oxyz, but also with respect
to the particles themselves. The exception is the rigid body for which
N
, = ro + M
rc = ro +rc 1 '~miri'
" ' (4.3.3)
i=1
-=-+
where ro = 00 denotes the position vector of the pole 0 which is the origin
of the axes Ox' y' z' fixed in the body, and rc
= OC denotes the position
vector of the centre of inertia C in the above axes which does not change
under the motion.
The concept of the tensor of inertia eO of the system of mass points is
more complicated.
The definition of a tensor of second rank is given in Sec. A.4. A dyadic
product ab of vectors a and b, also known as a dyad, is introduced in Sec.
A.2. A dyad is an example of a tensor of second rank since premultiply-
ing or postmultiplying the dyad ab by c yields vectors (c· ab or ab . c,
respectively) .
Operations over tensors are simplified by entering the dyadics of unit
vectors isik of the adopted system of coordinate axes. Denoting the com-
ponents of tensor P in these axes as Psk , this tensor can be represented as
a sum of the nine dyadics
3 3
P = LLPskisik, (4.3.4)
s=1k=1
because this notation conforms with the property of the tensor to yield a
new vector being multiplied by a vector
3 3 3 3 3
P .a = L L Pskisik·a = L is L Pskak= L iscs = c,
s=1 k=1 8=1 k=1 s=1
4.3 Tensor of inertia 159

see definition (A.4.8) of the tensor.


A simple example of the tensor is the unit tensor

1 0
o 1 (4.3.5)
o 0
Its dyadic form is given by

(4.3.6)

Another sort of tensor is a diagonal tensor

(4.3.7)

If also Ql = Q2, the diagonal tensor can be cast in the form

(4.3.8)

The following tensors of second rank

P x a and a x P (4.3.9)

are introduced in Sec. A.4. By means of the dyadic representation (4) of


the tensor we obtain
3 3 3
P X a = LLL PktasEtsrik i ,.., (4.3.10)
s=lk=lt=l

3 3 3

a x P = LLL PktasEskririt· (4.3.11)


s=lk=lt=l

It is easy to prove that the table of components of this tensor coincide with
the matrix, cf. (A.4.14).
Let us also notice the following formulae

(a x P) . b = a x p. b, (P x a) . b = p. (a x b). (4.3.12)

We consider now the dyadic representation of the tensor in a coordinate


system rotating with angular velocity w with respect to an inertial co-
ordinate system. Applying the formulae for differentiation of unit vectors
(2.7.6), we obtain

Psk [(w x i~) i~ + i~ (w x i~)l} =1> +w x P - P x w. (4.3.13)


160 4. Basic dynamic quantities

Here P* denotes a tensor whose components in the moving axes are equal
to the time-derivatives of the components of P in these axes. Formula (13)
is a generalisation of the rule of differentiating a vector relative to moving
axes on a tensor of second rank.
A tensor of the second rank is termed symmetric if

JDsk =JDks (k,s= 1,2,3). (4.3.14)

In the case of the symmetric tensor, the products a . P and P . a define


the same tensor. The expression
3 3
a .P .a =L L JDskasak (4.3.15)
s=l k=l

provides us with a representation of a quadratic form of projections of


vector a whose coefficients are determined by a symmetric tensor P.
We proceed now to consider the tensor of inertia of the material system
about point 0 which is the origin of the coordinate basis. Let e designate
the unit vector of axis ~A. The moment of inertia of the material system
about the axis is the sum of products of particle masses with the square of
the distance hi from the axis.
We have

e· ri = ri cos ai, hi = ri sinai, h~ = r~ - (e· ri)2 ,

where ri stands for the position vector of the point under consideration
and ai denotes the angle between r i and axis 0 A. The moment of inertia
about this axis is equal to
N N
JOA = Lmih~ = Lmi [r~ - (e.ri)2]. (4.3.16)
i=l i=l

The square of the scalar product e· ri can be cast as follows

(e· ri)2 = (e· ri) (ri . e) = e· riri· e. (4.3.17)

We can also write

e .E .e = 1, r~ = e . Er~ . e. (4.3.18)

Noticing that vector e can be placed beyond the summation sign in eq.
(16), we obtain the representation for the moment of inertia about the axis
N
JOA = e· L mi (r~E - riri) . e (4.3.19)
i=l
4.3 Tensor of inertia 161

as a quadratic form of projections of e which are the direction cosines of


the angles between axis OA and the coordinate axes. This quadratic form
is produced by symmetric tensor of second rank
N
8° = L mi (r;E - riri) (4.3.20)
i=1

which is named the tensor of inertia of the material system about point O.
The components of the tensor of inertia about point 0 relative to the
axes of the coordinate system with the origin at this point are, due to (20),
as follows

(4.3.21)

where
N N N
811 L mi (y; + z;), 812 = - L mixiYi, 813 = - L mizixi,
i=1 i=1 i=1
N N N
- '"""' m'x'Y'
~ t 1, 1" 8 22 =L mi (Z; + xl), 8 23 = - L miYizi,
i=1 i=1 i=1
N N N
- '"""'
~ m 'lXX·
Z z, 8 32 = - LmiYiZi, 8 33 = Lmi (x; +Y;),
i=1 i=1 i=1

cf. (A.2.21). The diagonal components represent the moments of inertia


about the axes
(4.3.22)
The non-diagonal components of the opposite sign are referred to as the
products of inertia and are denoted as follows
812 = 8 21 = -Jxy , 8 23 = 8 32 = -Jyz , 8 31 = 813 = -Jzx ' (4.3.23)
In terms of the introduced notation, the tensor of inertia is written in
the form

(4.3.24)

When a moving body is considered, the components of the tensor of


inertia change as the positions of the masses change relative to each other
and the coordinate system. The exception to this is the rigid body provided
that its tensor of inertia is calculated relative to the axes fixed in the body.
When a solid is considered, the sums in eqs. (1) and (21) are replaced by
definite integrals over the volume of the body.
162 4. Basic dynamic quantities

4.4 Transformation of the tensor of inertia


Given the tensor of inertia eO about point 0 we look for the tensor of
inertia eO' about point 0'. To this end, it is sufficient to set in eq. (3.20)

-----t
where 00' = ro denotes the position vector of the new origin 0' relative
---,---+
to the previous one and 0' Mi = r~ denotes the position vector of point M'
under consideration relative to 0'. Since

the following relationship is found


N N N

eO L mi (Er; - riri) + Er6 L mi + 2ro· L mir~E


i=1 i=1 i=1

or by eqs. (3.2) and (3.3)

e°=e0' +M (Ero-roro
2 , 1 (rero+rore
) +2M [ Ero ·re-"2 , , )] .
(4.4.1)

This relationship is considerably simplified when point 0' coincides with


the centre of inertia. Then r~ = 0 and

eO = ee+M (Er~ - rere) , (4.4.2)

where re is the position vector OC of the centre of inertia with respect to


the origin O.
Formula (3.20) provides us with such a definition of the tensor of inertia
about point 0 which is indifferent to a particular basis with the origin at
this point. Assuming the two sets of axes are parallel to each other and have
the origins at points 0 and 0' we find from eq. (1) the following formulae
defining the components of the tensor of inertia about point 0 in terms of
its components about point 0'

e~ e[ + M [(X61 + X62 + X63) Dik - XOiXOk] (4.4.3)

+2M [(XOlX~1 + X02X~2 + X03X~3) Dik - ~ (XOiX~k + X~iXOk)] .

Here XOl, X02, X03 denote the coordinates of point 0' in the coordinate sys-
tem OX1 X2X3 with the origin at point 0 and x~1' x~2' X~3 in the coordinate
4.4 'Transformation of the tensor of inertia 163

system O'x~x~x~ with the origin at point O. The axes OXI and O'x~ etc.
are assumed to be parallel. When 0' coincides with the centre of inertia C
of the system then XCi = 0 and XOi = XCi. Equation (3) takes the form

8~ = 8h, +M [(X~l + X~2 + X~3) t5 ik - XCiXCk] . (4.4.4)


For example,

8g = 8g + M (X~2 + X~3)' 8g = 8f2 - MXCIXC2· (4.4.5)


The first formula in the latter equation expresses the well-known Steiner
theorem on the moment of inertia about the parallel axis passing through
the centre of inertia.
We proceed now to constructing formulae for the components of the
tensor of inertia about point 0 with respect to two coordinate systems
0~1~2~3 and OXIX2X3 having the same origin. Let a denote the matrix
of direction cosines which brings the first system into coincidence to the
second one. Then

X = a~, ~ = a' x, (4.4.6)


where X and ~ are the column-matrices of a generic point M. The subscript
i is omitted in what follows.
We enter now the matrix of inertia about point O. Due to eq. (3.20) and
relationships (A.2.19) and (A.2.21) it can be presented by the formula

8=Lm(Ee~-~e). (4.4.7)

The elements of this matrix are related to axes O~l ~2~3. The components
of the tensor of inertia about the same point 0 with respect to the axes
OXIX2X3 produces a matrix denoted by 8* which is given by

8* = Lm (Ex'x - xx') (4.4.8)

or by virtue of eq. (6)

Since e is a scalar one can write


~

and moreover

as matrices a and a' coincide for all components of the sum. We then have
that
8* = a8a', 8 = a'8*a. (4.4.9)
164 4. Basic dynamic quantities

These are the formulae relating the matrices of inertia with respect to two
coordinate systems with the same origin. By the rule of matrix multiplica-
tion we obtain the formulae relating elements of these matrices which are
the components of the tensor of inertia
3 3
8;t = LL 8km(}:sk(}:tm' (4.4.10)
k=lm=l

For instance,

8r1 811(}:~1 + 822(}:~2 + 833(}:~3 +


28 12 (}:11(}:12 + 28 23 (}:12(}:13 + 2831 (}:13(}:11. (4.4.11)

The latter is a well-known formula expressing the moment of inertia about


a certain axis (in this case about OX1) in terms of the moments of inertia
and products of inertia in the adopted coordinate system Oe1e2e3 and
the direction cosines (}:11, (}:12, (}:13 of the angles between this axis and the
coordinate axes.
Because 8h is positive by definition for any real values of the variables
(}:11, (}:12, (}:13, the quadratic form obtained by means of matrix 8 is positive
definite, that is coefficients 8 sk satisfy the Sylvester inequalities

8 11 8 12 8 13
8 11 > 0, 8 12 822 823 > O. (4.4.12)
8 13 823 8 33
An expanded expression for the transformation of the products of inertia
is given by

8r2 8 11(}:11(}:21 + 822(}:12(}:22 + 8 33 (}:13(}:23 + 8 12 ((}:11(}:22 + (}:12(}:21)


+823 ((}:12(}:23 + (}:13(}:22) + 8 31 ((}:13(}:21 + (}:11(}:23). (4.4.13)

4.5 The principal axes of inertia


It follows from eq. (4.10) that

(4.5.1)

It is known that the directions of axes OX1X2X3 (in other words matrix (}:)
can be defined such that the inertia matrix about point 0 is diagonal

8i 0 0
8* = 0 82 0 (4.5.2)
o 0 83
4.5 The principal axes of inertia 165

These axes are referred to as the principal axes of inertia at point 0 and the
diagonal elements of matrix 8* are termed the principal moments of inertia
about this point, 8; being the moment of inertia about the principal axis
OX s . Equalities (1) for this axis (i.e. for a fixed s) yield the three equations
3 3
L a sk 8 kt = 8;ast = 8; L Dktask s = 1,2,3. (4.5.3)
k=l k=l
Omitting subscript s and assuming ask = f3k we can rewrite these equations
in the form
3
L f3 k (8 kt - Dkt 8*) =0 (t = 1,2,3) (4.5.4)
k=l
with
(4.5.5)
The determinant of system (4) should be equal to zero since, due to eq.
(5), not all values of f3 k are identically zero. This leads to the following
equation
8 11 - 8* 8 12 8 13
8 12 8 22 - 8* 8 23 = 0, (4.5.6)
8 13 8 23 8 33 - 8*
implying that the determinant of the following matrix
f(8*)=8-E8* (4.5.7)
vanishes. It can be proved that all three roots 8 s of this cubic equation are
real-valued since matrix 8 is symmetric.
It is necessary to recognize three cases:
a) The roots of equation (6) do not coincide. The deficiency of matrix (7)
is then equal to one, which means that there exists a non-zero minor deter-
minant of determinant (6) for any s = 1,2,3. Let the minor determinant
of the element 8 33 - 8* is non-zero, i.e.

where 8; is one of the roots. Under this condition the unknown variables
f3~ and f3~ can be expressed in terms of f3~, the latter being determined from
the normalisation condition (5). For s = 1,2,3 we obtain three directions
and form the matrix 11f3~ II

f3~ f3~ f3~


f3i f3~ f3~ (4.5.8)
f3i f3~ f3~
166 4. Basic dynamic quantities

In order to prove the mutual orthogonality of these directions we notice


that due to eq. (4)
3
L ,8%8kt = 8;,8: (t = 1,2,3),
k=l
which yields
3 3 3
L L,8%,8~8kt = 8; L,8:,8~
t=l k=l t=l
or by means of interchanging superscripts
3 3 3
L L,8k,8:8kt = 8; L,8~,8:'
t=l k=l t=l
The left-hand sides of both these equations are identical. This can be easily
proved by swapping the subscripts k and t and taking into account the fact
that 8kt = 8 t k. For this reason
3
(8; - 8;) L,8~,8: = 0,
t=l
and because of 8: =I- 8; we have
3
L,8~,8: = 0 (r =I- s).
t=l
The latter indicates that the directions corresponding to the principal axes
are orthogonal. Matrix (8) is the required rotation matrix Ct.
b) If equation (6) has two equal roots 8i = 8 2, then the deficiency
of matrix (7) for 8* = 8i = 8 2 is equal to two and there is a minor
determinant of the first order, say 811 - 8i = 811 - 8 2, which is not equal
to zero. For any triple of values ,8~ and ,8~ we have one equation (4) for
t = 1 and one equation (5). These values are also related to each other by
means of the equality

which expresses the orthogonality of the directions corresponding to the


first and the second axes of inertia. The direction of the third axis is deter-
mined uniquely as shown in a). The orthogonality of this axis to the first
and the second axes of inertia can be proved similarly. The non-uniqueness
of one of the values ,8~ and ,8~ indicates that one of the principal axes (the
first or the second) may have an arbitrary direction in the plane normal
4.6 Inertia ellipsoid 167

to the third principal axes. Matrix a describes an arbitrary rotation about


this axis.
c) All roots of eq. (6) coincide. Then all the minor determinants of the
first order of the determinant (6) vanish, i.e. all elements of determinant
(6) are identically equal to zero. Equations (4) are satisfied identically and
the nine elements of matrix (8) are related by six equalities expressing the
orthogonality of the trihedron of the principal axes. Matrix a remains un-
determined, that is any three mutually orthogonal directions can be taken
as directions of the principal axes of inertia at point O.
Provided that the directions of the principal axes at point 0 are given
by the unit vectors i;, the dyadic representation of the tensor of inertia at
this point takes the diagonal form

(4.5.9)

The orthonormalised trihedron ii, i2, ij is determined uniquely when the


principal moments of inertia are different. If 8i = 8 2 =1= 8j then the
trihedron is determined to be a rotation about ij. In this case, due to eq.
(3.8) the inertia tensor can be cast in the form

80 -- 8*E
1 + (8*3 - 8*) .*.*
1 13 13' (4.5.10)

Finally, when all three principal moments of inertia are equal, the inertia
tensor can be represented as a product of the principal moment of inertia
and the unit tensor

(4.5.11)

with any orthogonal trihedron being taken as the trihedron of the principal
axes.

4.6 Inertia ellipsoid


Construction of the inertia ellipsoid serves as an auxiliary means to illus-
trate the concept of the inertia tensor at point O. Let us consider a line
segment of length p

1
p=-- (4.6.1)
,jJOA
along the axis OA whose direction is given by the unit vector e. The coor-
dinates of point K which lies at the end of this segment are

a
X=--- (4.6.2)
,jJOA'
168 4. Basic dynamic quantities

where a, (3, '"Y designate the direction cosines of the angles between vector
e and the axes Oxyz. These are the direction cosines all, a12, a13 in eq.
(4.11) provided that e~l is replaced by JOA. Performing this replacement
and inserting a, (3, '"Y from eq. (2) we obtain, after some simplifications, that

(4.6.3)
This is the equation of a surface of the second order which is a locus
of end of the line segment (1). It follows from inequalities (4,12) that this
surface is an ellipsoid having the centre at point O. It is called the inertia
ellipsoid at this point. When the principal axes of inertia x*, y* z* are taken
as the coordinate axes x, y, z at this point then the equation for the ellipsoid
takes the form

(4.6.4)

Therefore, the axes of the inertia ellipsoid coincide with the principal axes
of inertia at point 0, the length of each axis, due to eq. (1), being given by

1 b=_l_ 1
a=-- c=--
J8I. (4.6.5)
J8f' ~'
The inertia ellipsoid at the centre of inertia C is referred to as the central
ellipsoid, the central axes of inertia at this point is termed the principal
central axes of inertia and the principal moments of inertia are named the
principal central moments of inertia. If a homogeneous rigid body has such
a shape that

then

er 2 e; 2 8; and a::; b ::; c. (4.6.6)


Thus, the central ellipsoid of inertia is extended along the axis Oz* and
compressed along the axis Ox* i.e. it is a certain "replica" of the body
form. For a = b = c the inertia ellipsoid transforms into a sphere which
corresponds to the body of a "cube-like" form.
The case e;; = 0 occurs for an infinitesimally thin rod, and the inertia
ellipsoid degenerates into a circular cylinder with the axis Oz*.
In passing we note that

1 1 1
i.e. 2" + 2' (4.6.7)
a ::; b2 c

the equality sign takes place only when 2: mx*2 = 0, i.e. when the body is
an infinitesimally thin plate lying in the plane Cy* z* .
4.6 Inertia ellipsoid 169

£,-i',
Ij.

FIGURE 4.1.

We assume that 0i # 0 and introduce into consideration


(4.6.8)

Inequalities (6) yields


El ~ 1, E2 < 1.
El -

In the plane E l , E2 they define a domain which is the interior of the rectan-
gular triangle OAB with OA = OB = 1, see Fig. 4.1. All possible ellipsoids
(not necessarily the inertia ellipsoids) for which a ~ b ~ c are mapped on
this domain. The border OA , i.e. E2 = 0, C = 00, describes elliptic cylinders
with axis Cz*, the border OB corresponds to the ellipsoids of revolution
about axis C x* flattened along this axis (a ~ b = c) , and AB represents
the ellipsoids of revolution about axis Cz* for which a = b ~ c.
Inequality (7) can be rewritten in the form
El + E2 2': 1.
One can see that the inertia ellipsoids correspond to the hatched domain
AC B which is a part of the domain for all possible ellipsoids. The border
AC describes the inertia ellipsoids of the bodies whose masses are dis-
tributed over a planar region in the plane Cy* z*. At point C we have
El = E2 = 0.5, i.e. 0i = 28 2 = 28 3, It is worth noting that the non-central
Kovalevskaya ellipsoid for which 8i = 8 2 = 28 3 is situated in at point
K. The ellipsoid corresponding to point L which is the point of intersection
of the bisectrix of angle B with the side AC is represented by
y2-1
E2 = y2 , l.e. 8; = 0.7078;' , 8 ; = 0.2938;'.
170 4. Basic dynamic quantities

This is the inertia ellipsoid which deviates most from the ellipsoid of revo-
lution.
In conclusion we notice that a similar, however not identical, graphical
construction of the domain of parameters defining the inertia ellipsoids is
suggested in [65].

4.7 The kinetic energy of rigid body


We consider next the case of a rigid body having a fixed point O. The
velocity of points in this body is determined by the kinematic formula
(2.7.8) with Vo = 0

Vi = W x r~. (4.7.1)
For this reason the kinetic energy is equal to
1 1
L '2 L
N N
T = '2 mivi . Vi = mi (w x r~) . (w x r~) =
i=l i=l
N N
~L mi W ' [r~ x (w x r~)] = ~ L mi [w 2 r? - (w· r~)2] .
i=l i=l

This equation is transformed using eq. (3.16)


w 2r,2
2
- (w . r')
2
(w . r')
1.
= w . (r'2E
'"
- r'r')
22
. w•
Using the definition of a tensor (3.20), we yield the following formula

T= ~w.eO.w (4.7.2)
2 '
where eO denotes the tensor of inertia of the rigid body at point O. The
kinetic energy is represented by a quadratic form of projections WI, W2, W3 of
the angular velocity vector w with the coefficients defined by the symmetric
tensor eO. The matrix form is as follows
1 ,
T= -we W ° (4.7.3)
2 '
where wand w' are a column-matrix and a row-matrix of projections of w
on the taken coordinate axes and eO is the matrix of inertia.
Particularly, if axes Ox' y' z' fixed in the body are taken as the coordinate
axes, then e~ are constants and the above form has constant coefficients.
An expanded form of eq. (2) or (3) is

T = ~ (egwi + e~2W~ + e~3W~ + 2egwlw2 + 2e~3W2W3 + 2e~IW3WI) .


(4.7.4)
4.7 The kinetic energy of rigid body 171

This equation simplifies when the principal axes of inertia at point ° are
used to

(4.7.5)

The above formulae express the kinetic energy in terms of the quasi-
velocities. When the generalised velocities {p, iJ, rp are used the expression
for the kinetic energy becomes cumbersome even for the case in which the
principal axes are taken

T ~ { [( 8i sin2 <p + 8; cos 2 <p) sin 2 19 + 8; cos219] {p 2+


(8i cos 2 <p + 8; sin2 <p) l + 8;rp2 + 28;{Prp cos 19 +
2 (8i - 8;) sin <p cos <p sin 19{PiJ } . (4.7.6)

We proceed now to the general case of a rigid body motion. By virtue of


the formula for velocities in a rigid body

Vi = Vo +w x r~

and thus

Vi· Vi = V5 + 2 (vo x w) . r~ + (w x r~) . (w x rD.


By using eq. (2) and definitions (3.2) and (3.1) we obtain

T = ~ [Mv5 + 2M (vo x w)· r~ + w· 8°· w]. (4.7.7)

Here Vo is the velocity of the pole 0, r~ denotes the position vector OC


of the centre of inertia of the body in the axes with the origin at the pole
0, and 8° denotes the inertia tensor at this point.
A simple expanded form of formula (7) is obtained when one uses the
axes fixed in the body. The projections VOl, V02, V03 of velocity of the pole
° are then introduced and for the principal axes of inertia one obtains the
kinetic energy as a quadratic form of six quasi-velocities

T = ~ {M (V51 + V52 + V53) + 2M [(V02W3 - V03W2) x~+


(V03WI - VOIW3) y~ + (VOlW2 - v02wd z~l + 8iwi + 8;w~ + 8;wD·
(4.7.8)

The above expression is drastically simplified when the centre of inertia


is taken as the pole. Then x~ = y~ = z~ = 0 and

T = ~M (Vbl + Vb2 + Vb3) + ~ (8iwi + 8;w~ + 8;w~) . (4.7.9)


172 4. Basic dynamic quantities

Another important particular case is that of rotation of a rigid body


about a fixed point 0, when the inertia ellipsoid at point 0 is an ellipsoid
of revolution with the axis defined by the unit vector e. By virtue of eqs.
(5.10) and (2) we obtain

T = ~ [8iw2 + (8; - 8i) (w· e)2] , (4.7.10)

where w . e is the projection of the angular velocity vector on the axis e, and
8 3 and 8i are the moments of inertia about this axis and perpendicular
to it, respectively.

4.8 Principal momentum and principal angular


momentum of a rigid body
It is known that the vector mivi equal to product of the particle mass
and its velocity determines the momentum of the particle. The principal
momentum of the system is denoted as Q and, by definition, equals
N N
Q = Lmivi = Lmiri. (4.8.1)
i=l i=l
By virtue of eq. (3.1)

(4.8.2)

where Vc stands for velocity of the centre of inertia, we obtain

Q=Mvc, (4.8.3)
that is the principal momentum of the system is equal to product of the
total mass and the velocity of the centre of inertia.
The principal angular momentum of the system of particles relative to
point 0, i.e. the geometric sum of the moments of momentum about this
point is given by
N
KG = Lmiri x Vi. (4.8.4)
i=l
Further on we will consider the motion of a system of particles not with
respect to the inertial axes Oxyz but with respect to axes Ox'y' z' whose
motion is prescribed by velocity Vo of its pole 0 and the angular velocity
w. Then, due to the theorem on addition of velocities, we have

(4.8.5)
4.8 Principal momentum and principal angular momentum of a rigid body 173

where rc=oG denotes the position vector of the centre of inertia with
respect to the above axes, Vc and Vc are the absolute and the relative
velocities of the centre of inertia, respectively, and

Qe=M(vo+wxrc), Qr=Mvc (4.8.6)


designate the vectors of the translational and relative momenta, respec-
tively.
The expression for the angular momentum can be represented in the
form
N
KG = L mi (ro + r~) x (vo + w x r~ + vD
i=l
N N
= Mrox(vo +w x rc)+Mrcxvox+ Lmir~x(w x r~)+ Lmir~xv~.
i=l i=l
We have
N N
Lmir~ x (w x r~) ~
~
m·2 (r'2w -
t
r'r~
'l'l
. w)
i=l i=l
N
L mi (Er? - r~rD . w = eO . w,
i=l
where eO denotes the inertia tensor at point O. Due to eq. (4)

(4.8.7)

represents the relative angular momentum about point 0 and we arrive at


the equality

(4.8.8)
For the case in which the system under consideration is a rigid body and
axes Ox'y'z' are fixed in it, we have v~ = 0 and thus

Qr =0, K? =0.
We then have

Q=M(vo+wxrc), (4.8.9)

KG = ro x Q + Mrc x Vo + eO. w, (4.8.10)


174 4. Basic dynamic quantities

where eO is the inertia tensor of the rigid body at the pole 0 of the axes
fixed in the body. In particular, if the centre of inertia is taken as this pole,
then rc = 0, ro = rc and

Q = Mvc, KG = rc x Mvc + eC . w. (4.8.11)

°
When the rigid body has a fixed point 0 and this point is assumed to
coincide with () then ro = 0, Va = and

Q = Mw x rc, KO = eO. w. (4.8.12)

The latter expression, being projected on the axes Ox' y' z' fixed in the
body, takes the form

Kf = 8 11 W l + 8 12 W2 + 8 13W 3, }
Kf] = 8 21 W l + 8 22 W 2 + 8 23 W 3, (4.8.13)
Kfl = 8 31 W l + 8 32 W 2 + 8 33 W 3,
since the products of inertia are equal to 8 st (8 -=I=- t) with the opposite
sign. When the principal axes of inertia at point 0 are taken as these axes
then

(4.8.14)

Returning to Sec. 4.3 it is worthwhile mentioning the special case of the


body rotating about a fixed point 0 in which the inertia ellipsoid is an
ellipsoid of revolution. Then

KO =8;'(w-w.ee)+8;w.ee, (4.8.15)

where e denotes the unit vector of the rotation axis of the ellipsoid of
revolution and the vector in parentheses is the projection of vector w on
the plane perpendicular to this axis.

4.9 The kinetic energy of a system under relative


motion
Let us consider the theorem on composition of velocities written in the
form

Vi = Va +w x r~ + v~,
where va and w denote velocity of the pole 0 and the angular velocity of
the moving axes, respectively, v~ is the velocity relative to these axes and
4.9 The kinetic energy of a system under relative motion 175

Vi is the absolute velocity of the particle. Using the well-known equality


(a x b) . c = a . (b xc) we obtain

v; v5 + 2 (va x w) . r~ + (w x r~) . (w x r~) +


2vo·v~+2w.(r~ xvD+v?

By means of eqs. (7.7), (8.6) and (8.7) we find

T ~ [Mv5 + 2M (va x w) . r~ + w . eO . wJ +
N

Va . Qr + w . Kr° + 2"1 '"""'


L..- miv i'2 . (4.9.1)
i=l

The first line of thi::; equation represents the kinetic energy Te of trans-
lational motion. Formally, this expression does not differ from the kinetic
energy which the system would have if its particles were fixed with respect
to moving axes Ox' y' z'. However it is necessary to bear in mind that in
this case the inertia tensor eO at point 0 does not remain constant under
a motion (which would be the case of a rigid body fixed with respect to
these axes). The term
N

Tr = ~ Lmiv? (4.9.2)
i=l

represents the kinetic energy of relative motion. Now we can write

T = Te + Va . Qr + w . K~ + Tr · (4.9.3)
Describing the position of the system's particles relative to the moving axes
by means of the generalised coordinates Q1, Q2, . .. ,Qn and assuming
(4.9.4)
we have

(4.9.5)

(4.9.6)

These quantities are linear in the generalised velocities and thus we can
introduce, as eq. (1.4) suggests, the part of the kinetic energy linear in the
generalised velocities
n

T1 = Va· Qr +w· K~ = LBsils, (4.9.7)


s=l
176 4. Basic dynamic quantities

where

(4.9.8)

The kinetic energy of relative motion is a homogeneous quadratic form of


the generalised velocities

(4.9.9)

Finally, the kinetic energy Te of the translational motion does not depend
on the generalised velocities ris and is understood as To in eq. (1.4).

4.10 Energy of accelerations


Construction of the differential equations of motion in the form suggested
by Appell assumes the new quantity

(4.10.1)

which is termed, by analogy with the kinetic energy, the energy of acceler-
ations. There is no need to derive an exact expression for S since Appell's
equations contains only derivatives with respect to the generalised accel-
erations iis. The terms which do not depend upon iis are immaterial and
can be omitted. In what follows S* stands for S with the immaterial terms
omitted.
We study first the case of stationary constraints. Due to eq. (1.3.9) we
have

S*

Using expression (1.2)

we find
4.10 Energy of accelerations 177

Considering permutations of subscripts r, s, k we obtain two other relation-


ships

Subtracting the first expression from the sum of the second and third ex-
pressions we have

(4.10.2)

The expression on the right-hand side is referred to as a Christoffel's symbol


of the first kind for matrix IIAsk II. The notation is [k, s; r] = [s, k; r] which
is known as Christoffel's square brackets, see (B.4.14)

(4.10.3)

Now we obtain
1 n n n n n
S* = 2L L Askiisiik +L LL [s, k; r] qsqkiir· (4.10.4)
s=lk=l r=ls=lk=l

Thus, given the expression for the kinetic energy T, the expression for S*
is easily obtained using eqs. (3) and (4), see [58].
In the case of non-stationary constraints, the summation in the first sum
over sand k and over r in the second sum is from 1 to n (as iin+1 = 0)
whereas the summation in the second sum over 8 and k is from 1 to n + 1.
The result is
n+ln+1 n n
LL [s, k; r] qsqk
s=l k=l s=lk=l
n
2 L [8, n + 1; r] qs + [n + 1, n + 1; r].
s=l
178 4. Basic dynamic quantities

Taking into account that

we obtain
n
2 L [8, n + 1; r] qs
s=l
[n + 1, n + 1; r]

Hence,

S* ~ tt
s=lk=l
Askiisiik + ttt
s=lk=lr=l
[8, k; r] qsqkqr + (4.10.6)

This is the expression for S* under nonstationary constraints. To construct


this expression it is sufficient to know only the coefficients of the expression
for T.
We proceed now to obtaining an expression for the energy of accelerations
in terms of quasi-accelerations ws. To this end, it is sufficient to derive the
equation for that part of S* which depends on the quasi-accelerations. Our
consideration will be restricted to the case of stationary constraints and
homogeneous dependences (1.5.1) of quasi-velocities upon the generalised
velocities. Taking time-derivative of formula (1.15) we find

In view of this we can write

S* (4.10.7)

a2 r· a2 r·
It should be stressed that a a' i=- a a' , see eq. (1.9.5). A further
7rk trs 7rs 7rk
calculation is carried out fully analogous to the above. By virtue of eq.
4.10 Energy of accelerations 179

(1.17) we have

oA;r
07rk = i=l mi
t ( ori o2ri
07r s · 07rk07rr
ori
+ 07rr
o2ri)
·07rk07rs '

oA;k
07r s
= t
i=l
mi (ori . 0 2ri
07r r 07r s07rk
+ ori. o2ri ),
07rk 07r s07r r
(4.lO.8)

oAks =
07rr i=l
t
mi (ori . o2ri
07rk 07rr07rs
+ ori. o2ri )
07rs 07rr 07r k '
Recalling the rule (1.5.17) of "differentiating with respect to the quasi-
coordinate" and introducing the generalised Christoffel symbols we have

bms ~Akr - bmr ~A;k) = [8, k; rl7r = [k, 8; rl7r. (4.lO.9)


uqm uqm
Subtracting now the third line in eq.(8) from the sum of the first and the
second lines yields

The difference of the "second derivatives"

n
1 " Irk
"26 I A*
sl
1=1
180 4. Basic dynamic quantities

and by analogy

Inserting this into eq. (10) and recalling that 'Y~k = -'Y~r yields

Equation (7) now takes the form

Application of this formula requires only expressions for the kinetic energy
T in terms of the quasi-velocities and the three-index symbols.
As an example let us consider the case of a rigid body rotating about
a fixed point O. Taking the principal axes of inertia at point 0 as the
coordinate axes we obtain due to eq. (7.5)

Hence, all expressions [8, k; rl7r = O. Using eq. (2.10.3) we obtain

S* ~ (8 1 wi + 82W~ + 83W~) (4.10.13)


+W1W2W3 (8 3 - 8 2) + W2W3W1 (8 1 - 8 3) + W3W1W2 (8 2 - 8 1 ).

Let us notice that the expression for S* in terms of the generalised coordi-
nates would be very cumbersome.

4.11 Energy of accelerations of a rigid body


An expression for the energy of accelerations, strictly speaking for S*, is
easy to obtain directly using formula for the acceleration of points in a rigid
body.
We start with the case of a rigid body having a fixed point 0 which is
taken as the origin of the basis axes fixed in the body. We have

Wi = .x,
W ri + W x (W x ri') .
4.11 Energy of accelerations of a rigid body 181

Retaining only the terms with w we obtain

The first term in the latter equation differs from the equation for the kinetic
energy T in that woccupies the place of w. Then, due to eq. (7.2)
N
"21 "~ mi (w. x r ')
i . (.w x r ') 1. eO.
i ="2w· - ·w. (4.11.2)
i=l

The component of the second sum in eq. (1) is transformed as follows

(w x r~) . [w x (w x rDl x r i') . ( WW· r ,i - W 2 r ')


( W• i
(w x w) .r~w . r~ = (w x w) . [r~ x (w x r~)l,

where, while manipulating, we took into account that

(w x rD· r~ = 0, (w x rD . w = (w x w) . r~, (w x w) . w = o.
The second sum in eq. (1) is then given by
N N
L mi (w x rD . [w x (w x rDJ (w X w) . L mir~ x (w x r~)
i=l i=l
(4.11.3)

where KO, due to eq. (8.4), is the principal angular moment of the rigid
body about the fixed point O. Using expression (8.12) for KO we arrive at
the equality

1. . eO.
S * = "2w .w . (w x KO) = "2w
+ W· 1. . eO. . (w x eO)
- . w + W· - ·W .
(4.11.4)

The last term also can be written in the form (w x w) . eO . w which is a


bilinear form of projections of the vectors w x wand w with coefficients
defined by tensor eO.
The matrix form of eq. (4) has the form

(4.11.5)

where w' is a row-column of projections of vector W, w is a column-matrix of


vector w, and w denotes the skew-symmetric matrix accompanying vector
w by means of rule (A.2.3).
182 4. Basic dynamic quantities

An expression for S* in terms of projections of the vectors wand wand


the components of tensor 8° is easily obtained using eqs. (4) and (5)

S* = ~ (8 n wi + e22W~ + 833W~ + 2812W1W2 + 2e23W2W3 + 2831W3Wl)


+ (W2W3 - W2W3) (8nWl + 812W2 + 8 13 W3) +
(W3W l - W3Wl) (8 21 Wl + 8 22 W 2 + e 23 W 3) +

(W1W2 - W1W2) (8 31 Wl + 832W2 + 833W3). (4.11.6)

Clearly, expression (10.13) is obtained if the axes fixed in the body coincide
with the principal axes of inertia.
Let us construct an expression for S* for the general case of motion of a
rigid body. In the equation for the acceleration of a generic point

Wi = Wo + w.x,
ri + w x (w x ri') (4.11.7)
the first and the second terms differ from the corresponding terms in the
formula for the velocity of a point in that Wo and w
replace Vo and w,
respectively. With this in view we obtain the corresponding terms of S*
by making these substitutions in expression (7.7) for the kinetic energy T.
Two terms in the expression for S* remain to be calculated. The first term
is the result of the scalar multiplication of the second and third terms in
(7) and is given by eq. (3) or by the second term in eq. (4). The second
term in the expression for S* is easily calculated
N
LmiwO· [w x (w x r~)l Mwo· [w x (w x r~)l
i=l
M (wo x w) . (w x r~).

Thus, in the general case of motion of a rigid body

S* ~ M w6 + M (wo x w) . r~ +M (wo x w) . (w x r~)


+2w. - . w + (w. x w ) . 80
1. 80' - . w. (4.11. 8)

A considerable simplification is achieved when the origin of the basis axes


is assumed to be the centre of inertia C. Then r~ = 0 and

S* = -1 M Wc
2
2 + -w·
1. e- C . w• + ( w• x w ) . eC
2
- . w. (4.11.9)

However this expression may have unnecessary terms, that is, the terms
which are independent of the accelerations. Such terms have partially been
omitted above. We recall now that due to the formula for the time-derivative
of the vector
• =Vo
Wo = Vo * +w x Vo,
4.12 Example calculations of the kinetic energy for multi-body systems 183

the projections of the vector ";0 on the axes fixed in the body are equal
to time-derivative of the projections vOs of vector Vo. This enables us to
*
replace in eq. (8) w5 by v5 +2 ";0 . (w x vo) and Wo x w by Vo x wand
to retain only the terms containing the quasi-accelerations, i.e. vectors ";0
and w. Simple manipulation yields

S* 1 *2
"2Mvo+M (vo+wxrc·
I) (*
voxw ) + (4.11.10)

* . +w x Vo ) . (.
M ( Vo w x rc 1. eO
I) +"2w. . (w. x w ) . eO
- . w+ - . w.

4.12 Example calculations of the kinetic energy for


multi-body systems
4·12.1 A gyroscope in Cardan's suspension
A description of the system and the notation used are given in Sec. 2.6.
The platform carrying the bearings of the outer gimbal is considered to be
fixed. The kinetic energy of the outer gimbal rotating with angular velocity
a about the axis O~ of the platform is

where Al denotes the moment of inertia of the outer gimbal about axis O~.
The angular velocity of the inner gimbal is described by the vector

W2 = ah + j)i2 = i~ a cos,6 + j)i; + i;a sin,6,


where i l is the unit vector of the axis of rotation of the outer gimbal, and
i~ are the unit vectors of the axes Ox*y* z* of the inner gimbal. Assuming
that the trihedron of these vectors defines the principal axes of inertia of
the gimbal we obtain

T2 ="21 ( A 2 .2 cos 2 ,6 + B 2 ,6·2 +


(Y
·2 sm
C 2 (Y
. 2 ,6 ) .

In order to derive an expression for the kinetic energy of the rotor we


write its angular velocity vector in the form

W3 = W2 + i; 0 = i~ a cos ,6 + j)i; + i; (a sin,6 + 0) .

Using expression (5.10) for the inertia tensor at point 0, we obtain

T3 ~W3 . [A3 E + (C3 - A 3) i;i;] . W3

~A3 (a 2 cos 2 ,6 + l) + ~C3 (0 + asin,6)2.


184 4. Basic dynamic quantities

The kinetic energy is thus equal to

1
T = '2 (AI + A2 COS 2 (3 + C2 sin 2 (3 + A3 cos 2 (3) a2

+~ (B2 + A3) /3 2 + ~C3 (p + asin (3)2. (4.12.1)

4.12.2 A shell carrying flywheels


A rigid shell rotating about a fixed point 0 is considered. Denoting the
principal moments of inertia at point 0 as A, B, C and the angular velocity
vector as w we have due to eq. (7.5)

To = '12 (AWl
2 + BW22 + CW32) .
The shell carries the axes of n flywheels. Let us first consider the simple
case of balanced flywheels. This means that the centre of inertia of each
flywheel lies on the rotation axis (i.e. any flywheel is balanced statically)
and this axis coincides with one of the principal central axes of inertia
(i.e. any flywheel is balanced dynamically). The unit vector of the rotation
axis of the i - th flywheel is designated by ei and is assumed to have
a constant direction with respect to the shell described by the direction
cosines ai, (3i' "Yi of the angles referring to the principal axes Oxyz of the
shell. The position vector of the centre of inertia Bi of the flywheel is
-=---->
denoted by ri = OBi' Finally, (Piei stands for the angular velocity vector
of the flywheel relative to the shell and the absolute angular velocity of the
flywheel is given by

Assuming that the equatorial moments of inertia are equal (Ai = B i ) and
denoting the axial moment of inertia by Ci , we obtain with the help of eqs.
(7.9) and (7.10) that

2Ti = mi Iw x ril 2 + (w + ei(Pi) . [AiE+ (Ci - Ai) eiei]' (w + ei(Pi)


= miw, (Eri . ri - riri) . w+Ai [W2 - (w· e i )2] + C i (w· ei + (Pi)2 .
(4.12.2)

The kinetic energy is obtained by addition of these expressions for all


flywheels with the kinetic energy of the shell, which results in the formula

(4.12.3)
4.12 Example calculations of the kinetic energy for multi-body systems 185

e

FIGURE 4.2.

where eO is the tensor with the components

e~ = A: i~ [Ai (f3~ + "d) + CrT + zi)],


mi }
(4.12.4)
eg = - 2: (A i a if3i + miXiYi)
i=1

etc.

4.12.3 The kinetic energy of a body carrying an unbalanced


flywheel
Let the flywheel axis described by the unit vector e pass through point 0
-=--+
whose position in the body is given by vector r =00. The unit vectors of
the trihedron of the principal axes of inertia at point 0 are denoted by
a, b, c and are shown in Fig. 4.2. The unit vectors el, e2 , e are fixed in
the shell. The angle t.p between the plane of the unit vectors b, c and the
plane e , e 1 describes the angle of rotation of the flywheel and A is the angle
between the principal axis of inertia c and the rotation axis. Th~osition
of the centre of inertia S of the flywheel is given by vector p =08 whose
projections on axes a, b,c are denoted by Cl,C2 , C3.
Let us consider eq. (9.1). The first group of terms in this formula describes
the kinetic energy of the translational motion. In this formula

and hence

m Iw x rl2 + 2m (w x r) . (w x p) +
Al (w· a)2 + Bl (w· b)2 + C 1 (w· C)2 . (4.12.5)
186 4. Basic dynamic quantities

The kinetic energy of the relative motion is determined by the following


expression

2Tr = (p2e. (Alaa + BIbb + Glee) . e = (Bl sin2 A + Gl cos2 A) (p2,


(4.12.6)

representing the kinetic energy of the body rotating about a fixed axis. As
the relative motion is considered, this fixed axis is e.
Now the terms in eq. (9.3) arising due to the translational and the relative
motion can be calculated. By means of eqs. (8.12) and (8.3) we have

K? = (Alaa + BIbb + Glee) . erp = (-bB l sin A+ eGl cos A) rp,

Thus,

(-Blw· bsinA + Glw· ecosA) rp +


m (w x r) . (e x p) (p. (4.12.7)

Adding eqs. (5), (6), (7) and the kinetic energy of the shell we find that

T = ~ [Awi + Bw~ + Gw~ + m Iw x rl2 + Al (w· a)2 +


Bl (w· b-rp sin A)2 + Gl (w· e-rpcosA)2] +
m (w x r) . [(w + erp) xp]. (4.12.8)

In order to make use of this equation we need the rotation matrices


g and k which make the axes Oxyz fixed in the shell coincident with the
trihedron el, e2, e3 and the trihedron of the principal axes of inertia a, b, e,
respectively. Their expressions are as follows

0!12 0!13 sin i.p - cos i.p o


g= 0!22 0!23 k= cos Acos i.p cos Asin <p -sinA
0!32 0!33 sin Acos i.p sin Asin i.p cos A

The rotation matrix which makes the trihedron of axes Oxyz coincident
with trihedron a, b, e is thus kg.
Let us introduce the row-matrices

1~ = I 1 0 0 II, 1~ = I 0 1 0 II, 1; = I 0 0 1 II· (4.12.9)


Then the row-matrices a', b', d of the projections of the vectors a, b, e on
axes Oxyz are given by

(4.12.10)
4.12 Example calculations of the kinetic energy for multi-body systems 187

and thus

w· a =l~kgw, w· b =l~kgw, w· c =l;kgw, (4.12.11)

where W denotes the column-matrix of the projections of w on the men-


tioned axes.
Let x denote the column-matrix of the projections x, y, Z of vector r on
axes Oxyz, p the column-matrix of the projections of vector p on these
axes and e the column-matrix of the projections e1, e2, e3 of vector p on
trihedron a, b, c. Then

e = kgp, p = g'k'e

and we have

}
Iw x rl2 = w· (Er· r - rr)· w = w' (Ex'x - xx')w,
(w x r) . [(w + ecp) xp] = w . (Ep· r - pr) . (w + ecp)
= w' (Ex'g'k'e - g'k'ex') (w + cpg'1 3).
(4.12.12)

The calculation is not continued, as an extended expression for the kinetic


energy given byeq. (8) would be very cumbersome. However the calculation
is somewhat simplified for the case when 9 = E and the rotation axis e is
parallel to axis Oz. Then

w· a =l~kw = WI sincp - W2 coscp,


W . b =l~kw = cos A (WI cos cp + W2 sin cp) - W3 sin A,
w . c =l~kw = sin A (WI cos <p + W2 sin <p) + W3 cos A.

Provided that the flywheel is only statically unbalanced, i.e. oX = 0 and


Al = B l , the expression for the kinetic energy takes the form

T = ~ [Awi + Bw~ + Cw~ + m Iw x rl2 + Al (wi + w~) + CdW3 + cp)2]


+ mw' (Ex'k'e - k'ex') (w + cp13). (4.12.13)
In this case we can assume that e2 = 0 without loss of generality. We
also assume that the axis of vector c is the principal axis of inertia at the
point of intersection with the orthogonal plane passing through the centre
of inertia of the flywheel. This point can be taken as the origin of trihedron
a, b, c and then e3 = 0, e = e l 1l and the calculation yields

mw' (Ex'k'e - k'ex') (w + cp13) = mel {cos<p [(W3 + cp) (W2Z - W3Y)-
WI (WlY - W2X)]- sin <p [W2 (WlY - W2X) - (W3 + cp) (W3X - WlZ)]).
188 4. Basic dynamic quantities

FIGURE 4.3.

It is clear that the same result can be obtained by extending the following
vector expression

[(w x r) x (w + ec,o) 1. p
and noticing that the projections of p on axes Oxyz are equal to 101 sin <p,
-101 cos <p, O.

4.12.4 A platform carrying gimbals with rotors


The bearing of the rotation axis of the outer ring of Cardan's suspension
are mounted on the rigid body So whose motion is prescribed. The inner
ring whose axis is mounted on the outer ring is a platform used to carry
gyros which are rotating rotors placed in the gimbals, the latter being able
to rotate relative to the platform, see Fig. 4.3.
The rotation axes of the outer and inner rings are assumed to be perpen-
dicular to each other and intersect at the centre of inertia of the outer ring
C. The velocity of the centre of inertia caused by the prescribed motion of
the body So is denoted by V and the angular velocity vector of the body
So is denoted by n.
The orthogonal trihedron is of the unit vectors is bound to the outer
ring, such that hand h coincide with the rotation axes of the outer and
inner rings, respectively. The axes of the coordinate system Cx'y' z' bound
to the platform are given by the unit vectors i~ with i~ = i2 . The bearings
of the gimbals Kk are fixed to the platform and ak, bk, Ck denote the tri-
hedron of the orthogonal unit vectors bound to the gimbals K k , vectors ak
and Ck coinciding with the rotation axes of the gimbal and the rotor Rk,
respectively. The angle of rotation of gimbal K k about axis ak is denoted by
4.12 Example calculations of the kinetic energy for multi-body systems 189

'Yk and the position of the point Ck of intersection of the axes of rotor Rk
and gimbal Kk is given on the platform by the position vector CC~ = r~.
This point is the centre of inertia of rotor Rk whereas the centre of inertia
of gimbal Kk, along with a balance mass, is offset from Ck in the direction
of axis Ck.
The kinetic energy of the system is composed of the kinetic energies of
the outer ring, the platform, the gimbals and the rotors.
a) The outer ring. The velocity of its centre of inertia is V and its angular
velocity is 0 + ha, where a denotes the angle of rotation about axis i 1 .
Then

Tl = ~Ml V2 + ~ (0 + ha) . sf . (0 + ila).


b) The platform. Its angular velocity is equal to

Wp = 0 + ha + i~t3, (4.12.14)
where f3 designates the angle of rotation of the platform with respect to the
outer ring. The position vector of the centre of inertia C of the platform is
denoted r*. Due to eq. (7.7) the kinetic energy is given by

T2 = ~ [M2V2 + 2M2V· (wp x r*) +Wp· sf . wp] .

c) The gimbal K k • Its angular velocity and the velocity of point Ck are
given by w p + ak''tk and V + w p x r~, respectively. The position vector of
the centre of inertia Sk is CkS~ = CkCk. In accordance with eq. (7.7) the
kinetic energy Tk of the gimbal K k is as follows

Tk ~ {m~ IV + Wp x r~12 + 2m~ (V + Wp x rU·


[(wp + ak''Yk) x ckCk] + (wp + ak'Yk) . S,Ck. (wp + aki'k)}.
Here S'Ck denotes the inertia tensor at point Ck. Assuming equality of the
equatorial moments of inertia A~ and Bi. about axes ak and bk we have
SiCk = A~E + (Ck - A~) CkCk.

d) The rotor Rk. Its angular velocity is Wp + ak'Yk + Ck<Pk, where <Pk
denotes the angular velocity of rotor Rk. As the centre of inertia coincides
with point Ck we obtain

T k" '12 {mk"IV +wp x rk 12 +


I

(wp + ak''Yk + Ck<Pk) . S"c k • (wp + ak''Yk + Ck<Pk)] ,


where
190 4. Basic dynamic quantities

denotes the inertia tensor at point Ck'


We proceed now to the expression for the kinetic energy of the platform
and the bodies mounted on it. It has the term which is proportional to V2

"21 MV 2 ="21 [ M2 +~
~( ' + mk
mk II)] V.
2

The term with the factor V x Wp is given by

(Vxwp)· [M2r*+ ~(m~+m~)r~+ ~m~EkCk]


By properly mounting the balance masses on the platform, one can de-
termine the centres of inertia of the platform, the rotors and the gimbals
coincident with point C, i.e.
n
M2r* + L (m~ + m~) r~ = O.
k=l
In this case the expression for the term associated with V x Wp is as follows
n
(V x wp)· Lm~Ekck'
k=l
The terms quadratic with respect to W pare

~W P . { ef + ~ (m~ + mD (Er~ . r~ - r~rU +


~ [EAk + (Ck - Ad CkCk] + ~ m~Ek (Er~. Ck - r~ck) } . Wp.
Here Ak = A~ + A~ and C k = C k + C k. Let us denote the constant part of
the tensor in the braces as Q. The latter equation then takes the form

~wp . Q. wp + ~wp . t [(Ck - A k ) CkCk + m~Ek (Er~ . Ck - r~Ck)J . Wp.


k=l
The terms depending upon the angular velocities '"r k of the gimbals with
respect to the platform are given by

- tm~Ek (V + Wp x rU· bk'"rk + tAk'"rkWP ' ak + ~ tAk'"r~.


k=l k=l k=l
Finally, the terms associated with the angular velocities of the rotors are
4.12 Example calculations of the kinetic energy for multi-body systems 191

. - d~,

o.g.

FIGURE 4.4.

where H k = Cr
I{; k denotes the " proper kinetic moment of the rotor" .
Along with the kinetic energy of the ring we arrive at the following
expression for the kinetic energy of the system

T = ~ (M1 + M) V2+~ (0 + i 1 a) ·ef ·(0 + i 1 a)+(V x wp).:t m~Ekck


k=l

+ ~wp . { Q + ~ [(Ck - A k ) CkCk + m~Ek (Er~ . Ck - r~Ck)J } . wp


+ :t m~Ek
k=l
(V + wp x rD . bk'h + ~ :t
k=l
Ak'Ykwp . ak

1 n n ( 1 H 2)
+2 LAk'Y%+L HkWP'Ck+2C~ (4.12.15)
k=l k=l k

4.12.5 Gyrovertical
As an example of application of equation (15) we consider a gyrovertical
schematically depicted in Fig. 4.4 for the initial position of the platforms
and the gimbals, [66]. The base is assumed to be fixed , i.e. V = 0 and 0 = O.
The rotation axes of the gimbals passing through points C 1 , C 2 , C 3 , C 4 are
perpendicular to the plane i~, i3 of the platform and their unit vectors are

The gimbals are arranged in pairs so that 1'1 = 1'2 and 1'3 = 1'4' This
effect can be reached either by means of a gear train or (for small angles)
192 4. Basic dynamic quantities

an anti parallel link mechanism as shown in the above Figure. The rotors
in the linked gimbals rotate in opposite directions (counterclockwise being
observed from the end of vectors Ck). The position vectors of points Gk are

r~ = ai~ + bi~ + ci~, r~ = -ai~ + bi~ + ci~,


r~ = ai~ - bi~ + ci~, r~ = -ai~ - bi~ + ci~.

The masses m' and the inertia moments of all gimbals A', G' and all the ro-
tors (mil, A", Gil) are assumed to coincide. The balance masses are absent.
The angular velocity of the platform is given by

wp = i~acos,8 + i~,8 + i~asin,8.

Expressions for the relevant vectors are given by

C1 =i~sin'Y1-i~cos'Y1' C2 =i~sin'Y1 +i~COS'Y1'


C3 = -i1 COS'Y3 - i~ sin'Y3, C4 = i~ COS'Y3 - i~ sin'Y3'

Formula (15) for the kinetic energy is considerably simplified and takes
the form
4
T 1 11 a·2
-8 + -Wp
1 .8 C2 . Wp + -mwp'
1 L( '"
Erk . rk - ') . Wp +
rkrk
2 2 2
k=1
4
2A (a 2 + ,82) + +~ (G' - A) L (wp . Ck)2 + A bi + 'Y~) +
k=1
4 4
Awp . L ak'Yk + ~GII L (0k + wp' Ck)2 . (4.12.16)
k=1 k=1
Assuming knowledge of angles a,,8, 'Y1' 'Y3 and taking into account the
above expressions for wp,r~ and Ck we obtain

T = ~ [8 11 + 8 21 + 2A + 2G' + 4m (b 2 + c2)] a 2 + ~ [822 + 2A + 2G'

+4m (c 2 + a2)],82 + A bi + 'Y~) + ~GII [(01 + a'Y1 -,8f +


(02+ a 'Y1 +,8f + (03- a -,8'Y3f + (04+ a -,8'Y3f]· (4.12.17)

Here 8 11 denotes the moment of inertia of the outer ring about its rotation
axis, 8 21 , 8 22 denote the moments of inertia of the platform at point G
about i~ and i~, respectively, m and A are the total mass and the total
equatorial moment of inertia of the gimbal and the rotor, respectively.
4.13 Examples of kinetic energy and energy of accelerations 193

4.13 Examples of kinetic energy and energy of


accelerations
We now direct our attention to the examples of non-holonomic systems
studied in Sees. 2.10 and 2.11. It will be shown that when constructing
the equations of motion for a system with non-holonomic constraints one
needs two expressions for the kinetic energy, namely with and without the
non-holonomic constraints. For obtaining the energy of accelerations the
non-holonomic constraints should be taken into account.

4.13.1 Rolling sphere


Assuming that the centre of inertia of the sphere coincides with its geo-
metric centre we have due to (7.7)

T = "2IM(·2
Xo + Yo.2) +"218(-2 -2 + W3
- WI + W2 -2) , (4.13.1)

where 8 denotes the moment of inertia of the sphere about an axis passing
through its centre. For a solid sphere
2
8 = -Ma 2
5 .

Using expressions (2.10.19) for the quasi-velocities W4 and W5 we have

T ="2IM[72(_2 -2)
5"a WI +W2 2 2W3
+ 5"a - 2+2 aw2W4
-- - 2 - - +W4
aWIW5 -2 +W5
-2] .

(4.13.2)

In order to derive an equation for the energy of accelerations we turn to


eq. (11.9). For the sphere

2 2
8·w=-Maw.
5
The directions of the vectors 8 . wand ware seen to coincide, hence

(wxw)·8·w=0

and the latter term in eq. (11.9) vanishes. By virtue of eq. (2.10.19) we
have
•• ••
Xo =W4 +a W2, Yo =W5 -a WI .

Therefore
194 4. Basic dynamic quantities

The term ~w. e . w in eq. (11.9) differs from the corresponding expression
for the kinetic energy in that w replaces w. Hence, while llsing projections
of the vector on the fixed axes (which is the case) or on the axes fixed in
the body it suffices to replace W2 by W2 to obtain

We have

S* = T(w), (4.13.3)

which means that in order to construct the expression for the energy of ac-
celerations it is sufficient to replace the quasi-velocities Ws in the expression

for the kinetic energy by the quasi-accelerations ws.

4.13.2 Rolling ring


To calculate the kinetic energy we turn to eq. (7.7). Assuming that the
centre of inertia coincides with the geometric centre we have rc
= O. Using
eqs. (2.10.25) and (2.10.26) we obtain

The moments of inertia of an infinitesimal thin ring about the axis per-
pendicular to its plane and an arbitrary axis in the plane are respectively
equal to Ma 2 and ~Ma2, both axes passing through the centre O. For this
reason, the tensor of inertia at point 0 is

(4.13.5)

Recalling formula (2.10.13) for vector wand notation (2.10.14) we arrive


at

Hence

Calculation of the energy of accelerations is more complex. In order to


apply formula (11.9) we need equations for the acceleration of the centre
4.13 Examples of kinetic energy and energy of accelerations 195

of ring and the angular acceleration. We start with the acceleration of


the centre of ring, taking into account the presence of the non-holonomic
constraints given by eq. (2.10.24). Using notation (2.10.17) we obtain the
following representation for the velocity vector in the "half-fixed" system
n, nl' i3

Due to eq. (2.9.7) and taking into account notation (2.10.14) we obtain

• W2 db =0
nl = - - - n
sin {J , dt
and thus

While calculating w6 we drop the terms which do not contain ws , such that

(4.13.7)

We consider the derivation of equations for the angular velocity vector.


By virtue of eq. (2.10.13) we have

(4.13.8)

and using formulae (2.9.8) we obtain

W= [WI + W2 (W3 - W2 cot {J)] n + [W2 - WI (W3 - W2 cot {J)] n ' + w3i~.
(4.13.9)

To find the latter term in eq. (11.9) it is sufficient to take into account only
the terms that include ws. Up to this order of accuracy we have

Further on,

and hence

(4.13.10)
196 4. Basic dynamic quantities

We represent vector w in the form

W =W* +e, e = W2 (W3 - W2 cot 19) n-WI (W3 - W2 cot 19) n',

where c:, denotes the vector whose projections on the axes of the "half-
bounded" trihedron are respectively WI, W2, W3. Discarding the terms which
do not depend on Ws we have

I ' .e
-w - .'
w = -l *
we. - .*
W +e . e*
-. W
2 2
= 21 M a 2 [(.2 1.2
w3 + 2WI 1. 2)
+ 2W2 + (w3 - W2 co t u.Q) (WIW2
. . )] .
- WIW2

(4.13.11)

Multiplying the first expression by ~ M and adding the result to eqs. (10)
and (11) we obtain

S* 21 M a 2 [3 ·2
2WI 1.2 +
+ 2W2 2.w32 + 2W2 (WIW3
. -
. )
wIw3 +

2W3 (WIW2 - WIW2) - w2 (WIW2 - WIW2) cot 19]. (4.13.12)

4·13.3 Two-axle trolley


While calculating the kinetic energy with discarded non-holonomic con-
straints we should consider the components of the velocities of points A
and B as well as the velocities of contact of the wheel with the road as
being non-zero. In what follows these components are designated by low-
case letters, while capital letters correspond to the velocities of the motion
under constraints.
The trolley described in detail in Sec. 1.10 consists of six bodies, namely
the rear and front axles and four wheels. We will need the velocities of
points A, C, B and the wheel centres. Using the notation of Sec. 1.10 and
the directions of vectors il,i2,i~,i~ shown in Fig. 1.3, we have

VA = il (-WI sinx+w7cosX) +i2W2, V A = iIW7COSX,


VB = i~W7 + i~WI' VB = i~W7.

Let UI (and correspondingly Ud denote the angular velocity of the rear


axle which is, due to Sec. 1.10, given by

'.0
UI = 13 1(
U = T WI cos X + W7
.
sm X -) .
W2 13, (4.13.13)

where h is the unit vector perpendicular to the road plane.


4.13 Examples of kinetic energy and energy of accelerations 197

The velocities of the wheel centres of the rear axle are given by

VI VA + Ul x i 2 a = v A - ila~,
V1 · a .
V A-IlTW7S111X=W7 cosX-Tsmx ( a.). II,

v2 VA - Ul x i 2 a = v A + ila~,

V2 · a .
V A+IlTW7S111X=W7 ( cosX+T smx a.). II,

respectively. The angular velocities having subscripts 3 and 4 are composed


of the angular velocity of the axle Ul (and correspondingly U r) and the
angular velocities of rotation about the axle which are

. . = rl
12YI 1 [- ( .
SIll X +
acos )
T X WI + a
TW2 - W3 + (cosx a.)].
smx -T W7 12,

(4.13.14)

. . = rl
12Y2 1 [- ( .
SIll X -T
acos ) X
a
WI - TW2 - W4 + (cos X + a.)].
T SIll X W7 12,

(4.13.15)

provided that the constraints are discarded. When the constraints are taken
into account then WI = W2 = W3 = W4 = 0, and we have

with VI and V2 being the absolute values of vectors VI and V 2.


The angular velocity of the axis of the front axle is

(~+ X) i3 = ~ (WI cos X + W7 sin X - W2 + lws) i3,


1
T (W7 sin X + lws) i3 ,

where Ws = X. The velocities of the wheel centres are determined by the


following equalities

V3 VB+U2Xi~c=VB-C(~+x)i~,
V3 [W7 (1 - Ysin X) - cws] i~,
V4 VB-U2Xi~c=VB+C(~+X)i~,
V4 [W7 (1 + Ysin X) + cws] i~,
198 4. Basic dynamic quantities

whereas their angular velocities are as follows

U ., V3
5 = U 2+ 12-'
T2

U6 = U2 ., V4
+ 12-·
T2

Here
(4.13.16)

(4.13.17)

Now we proceed to construct an expression for the kinetic energy. Let


Ml and M2 denote the masses of the rear and the front axles (without
wheels) and 8 1 and 8 2 denote their moments of inertia about axes which
are perpendicular to the road and pass through points A and B. The masses
of the rear and the front axles are denoted by ml and m2, respectively. The
tensors of inertia of the wheels at their centres are respectively equal to

J~ (i1i1 + i3i3) + J1hh, J~ (i~i~ + i3ia) + Jai~i~.

Now applying formulae (7.7) and (7.8) we obtain the following expression
for the kinetic energy of the rear axle
1 2 , 1 ·2
Tl = 2M1VA + Ml (VA X Ul)· rc + 281'!9 ,

where r~ = i1s = ilAC. The kinetic energy of the wheels of the rear axle is

1 (2 2) 1 , ·2 1 ( . 2 . 2)
T3 + T4 = 2m1 VI + V2 + 22J1'!9 + 2J1 CPl + CP2 .
The kinetic energy of the front axle is

T2 =
1 2 1 (.
2M2VB + 282 '!9+ws
)2
whereas the kinetic energy of the wheels of the front axle is

1 (2 2) 1 ,(. )2 1 (.2 .2)


T5 + T6 = 2m3 V3 + V4 + 2 2J3 '!9 + Ws + 2Ja CP3 + CP4 .
Collecting the above expressions we obtain the kinetic energy of the
trolley

2T = [ 2 2]
(Ml + 2ml) (-WI sin X + W7COSX) +w2 + 2M1W2'!9S + .
(81 + 2mla 2 + 2JD ~2 + (M2 + 2m2) (wi + w~) + (4.13.18)

(82 + 2m3c2 + 2J~) (~+ ws) 2 + J 1 (<pi + <p~) + Ja (<p~ + <p~) ,


4.13 Examples of kinetic energy and energy of accelerations 199

where 19 and CPs are given byeqs. (13)-(17).


If we take into account the non-holonomic constraints we obtain the
following expression

2T = (J.L + J.Ll sin 2 X) w~ + Zvw~ + 2VWSW7 sin X, (4.13.19)

where

v = (4.13.20)

As mentioned above, the constrains are taken into account when cal-
culating the energy of accelerations. For this reason we use expressions
for the velocities and the angular velocities of the actual motion. While
differentiating one should bear in mind the relationships

dit _ V • _. W7 sin X dh V . . w7 sin X


-d-t - 1 X 11 - 12--Z-~ -d-t = 1 X 12 = -1 1 - -Z----'...;. (4.13.21)

di~ -_
dt V ./ _./ (W7 sin x
+ Ws ) di2 __ ./ (W7sinx )
1 X 11 - 12 Z , dt - 11 Z + Ws .
(4.12.22)

We obtain the acceleration of point A and the angular acceleration of the


rear axle
2
WA= VA = it (W7COSX - W7WS sin X) + h ~7 sin X cos X,

. 1
VI = hy (W7 sin X + W7WSCOSX) ;
the acceleration of the wheel centres and the angular accelerations of the
wheels

WI = VI = il [W7 (cos X - Tsin X) - W7WS (sin X + Tcos X)] +


.w?(
12T a.). 'W 'W
COSX-ySlllX SlllX=11 11+ 12 12,
200 4. Basic dynamic quantities

For the front axle

WB =
. 11./ W7. - 12./ (w?T sm. X + W7WS) ,
YB =

W 3 = Y· 3 = 11
./ [.W7 (1 - Y
c.
sm X) - yW7
c WS cos X - CWs
. ] +

W 4 = V 4 = i~ [W7 (1 + Ysin x) + yW7 w s cos X + cLuS] +

i; (~7 sinx+ws) [W7 (1 + ySin x ) +cws] = W41,i~ + W42,i;,

. . 1 / /
U 5 = U 2 + 2' (i2 W 31 , - il W 32, ) ,
r
The energy of acceleration of the rear axle is obtained from eq. (11.8) since
the centre of inertia C does not coincide with pole A. Equation (11.9) is used
for the front axle and the wheels. In the planar motion under consideration
eq. (11.8) is written in the form

s; = ~MIW1 +M1WA' (VI X Sil - U;sh) + ~811V112


Only terms containing W7 and ws should be taken into account. Thus, the
second term in the latter equation drops out and we have

S 1* .2 1 (8 1
='21 M lW7+'2 [2- M 1 ) (.2 . 2 X+ 2'W7w7wSsmxcosx·
w7 sm . )

Calculation for the front axle yields

. 2 1 8 2 (. 2 sm
. 2 l2 . 2
S 2* '12 M lW7+'2[2 w7
2. .
X+ W7w7wSsmxcosx+ ws+
2lwsW7WS cos X + 2lwsW7 sin X) .
4.13 Examples of kinetic energy and energy of accelerations 201

We have for the rear wheels that

S3* + S4* = "21 { ( 2ml + 2Jl) .


ri W7+ 2

{ [2l~{ - (1 - ~:) (2ml + 2:rl )] (w~ sin2 X + 2W7W7WS sinx cos X) } ,


and for the front wheels that

Adding these expressions and applying notation (20) we obtain

2S* (IL + ILl sin 2 X) W~ + lvw~ + 2VW7WS sin X cos X +


2ILl W7W7WS sin X cos X + 2VWSW7WS cos X. (4.13.23)

The first three terms can be obtained by replacing quasi-velocities W7


and Ws in the expression for the kinetic energy with quasi-acceleration W7
and ws, respectively. All other terms are given in eq. (10.12).
These additional terms can be split into two groups. The first group
denoted by Rl contains Christoffel's square brackets. In the case under
consideration the coefficients in the expression for the kinetic energy (18)
depend upon only one generalised coordinate q4 = X, by virtue of eq.
(1.10.14) only b4S = 0, while the other elements of the fourth row of matrix
b are identically equal to zero. Formula (10.9) takes the form

[s, k; rl7r = ~ :X (b4k A;r + b4s A kr - b4r A;k) ,

where only those square brackets do not vanish which have at least one
index 8, i.e.

. _ loA kr [ loA;r
[8,k,rl7r - "28X' 1
s,8;r 7r = "2ax'
[ k.8l = _.!. OA;k [
8,8;r 17r =
oASr
ax' (4.13.24)
s, '7r 2 oX '
loAss
[s, 8; 8l7r = [8, s; 8l7r = 0, [8,8; 8l7r ="2 oX .
Now we can write the first group of terms

Rl W7 (w~ [7, 7; 7l7r + 2W7WS [7,8; 7l7r + w~ [8,8; 7l7r) +


+ws (w~ [7, 7; 8l7r + 2W7WS [7,8; 8l7r + w~ [8,8; 8l 7r ) .
202 4. Basic dynamic quantities

By virtue of eq. (19)

A*77 = J-t
. 2
+ J-tl sm X, A~s = v sin X, Ass = lv,

and eq. (24) we can express Rl as

Rl = W7 (2J-tlw7wssinxcosx + vw~COsX) - J-tlwsw~sinxcosx. (4.13.25)

We proceed now to calculating the second group of additional terms in


eq. (10.12)

R2 = LLLLWrWsWkAkz'Y~r = W7 WS LLAkz'Y~7Wk +
r ski k I

WSW7 L L AkIWk/~S = (WSW7 - W7WS) L AklWk.


k I k

Here equalities (10.12) were used. The values of Ah and ASl are obtained
from eq. (18) for the kinetic energy

A~l = J-tl sin X cos x, ASl = v cos x.

Then we have

(4.13.26)

and finally

(4.13.27)

Thus expression (23) is obtained. This calculation was performed in or-


der to demonstrate that it is possible to derive an expression for the energy
of accelerations without knowledge of the acceleration of the system's par-
ticles. To achieve this it is sufficient to possess expressions for the kinetic
energy (found under the condition of discarded constraints) and the three-
index symbols. The required result is achieved by means of the calculation
due to eq. (10.12).
5
Work and potential energy

5.1 Generalised forces


The sum of the elementary work due to the application of forces Fi at
points Mi of a system undergoing virtual displacements 8r i of these points
from their positions at a fixed time instant is given by
N

8'W = LFi' 8ri. (5.1.1)


i=l

The notation 8' is used to indicate that we are dealing with an infinitesimal
quantity which is not a variation of the quantity W.
Replacing in eq. (1) the virtual displacements by expressions in terms of
variations of the generalised coordinates 8qs we obtain

(5.1.2)

The quantity

(5.1.3)

is referred to as the generalised force corresponding to the generalised co-


ordinate qs. Due to eq. (2) an expression for the elementary work in terms
204 5. Work and potential energy

of the generalised forces is given by


n
8'W = LQ88q8. (5.1.4)
8=1

Hence, the generalised force Q8 is equal to the coefficient of variation 8qs


of the corresponding generalised coordinate in the expression for the ele-
mentary work of the applied forces in the infinitesimal displacement from
the position under consideration.
Imagining that the points of the system are subject to a general virtual
displacement, i.e. a displacement such that all the generalised coordinates
are varied simultaneously, we find the all n generalised forces are given by
eq. (4). Provided that only a generalised force Qk is required with k fixed
it suffices to separate such a virtual displacement from the possible ones
so that only generalised coordinate qk is varied while the others are kept
constant. The elementary work (8'W)k' due to eq. (4), is proportional to
8qk

(5.1.5)

and the generalised force Qk is determined as the factor of 8qk in the latter
expression. It is good practice to calculate all Qk in this way.
When using quasi-coordinates we determine the virtual displacement 8ri
by means of eq. (1.6.14). Consequently the elementary work is represented
in the form

(5.1.6)

The quantity

(5.1.7)

is termed the generalised force corresponding to the quasi-coordinate 7r 8.


Using eq. (1.5.17) we can rearrange the latter equation, to obtain

(5.1.8)

The inverse relationships are as follows


n
Qs = LarsPr, (5.1.9)
r=1

where ars denotes the elements of matrix (1.5.7).


5.2 Elementary work of forces acting on a rigid body 205

Provided that the relative motion of a system of material particles is


studied, then the position of point Mi referring to the inertial axes is de-
termined in general by the equation

(5.1.10)

where ro is a prescribed function of time. The virtual displacement is, by


definition, or~, the variations of the vectors ri and r~ coincide, and the
expression for the generalised force takes the form

(5.1.11 )

5.2 Elementary work of forces acting on a rigid


body
Let us consider a free rigid body and denote the forces acting on it by
F 1, F 2 ... , F N. The virtual displacement of point Mi in the body is given by
the relationship

ori = oro + 0 x r~, (5.2.1)

in which oro denotes a virtual displacement of the origin (the pole) of the
axes Ox' y' z' fixed in the body, 0 the vector of infinitesimal rotation of the
body, and r~ = (5J;/;, the position vector of the point in question.
Therefore,

N N N

o'W LFi' 8ri = LFi' oro + LFi' (0 x r~)


i=l i=l i=l
N N
oro' L F i + 0 . L r~ x F i· (5.2.2)
i=l i=l
Since

(5.2.3)

is the resultant force of the system of forces under consideration and

N
Lr~ x Fi = rnG (5.2.4)
i=l
206 5. Work and potential energy

is the principal moment about the pole 0, we obtain the expression for the
elementary work in the form

b"W = V . b'ro + rno . o. (5.2.5)

If we take the coordinates xo, Yo, Zo of the pole 0 with respect to the fixed
axes Oxyz and the Euler's angles 'ljJ, {}, t.p as the generalised coordinates,
then

(5.2.6)

Here is and i~ denote the unit vectors of the fixed and moving axes, respec-
tively, and n stands for the unit vector of the nodal axis. Relationship (5)
reduces then to the form

b"W = l/j8xo + V2b'yO + V3b'zo + rno . i3b''ljJ + rno . nb'{} + rno . i;b't.p.
(5.2.7)

Due to eq. (1.4) we have

Ql = VI, Q2 = V2, Q3 = V3, (5.2.8)

i.e. the generalised forces, corresponding to the coordinates xo, Yo, Zo, are
the projections of the principal vector on axes of the fixed coordinate system
Oxyz. Furthermore

Q4 = rno . i3 = m3, Q5 = rno . n = mN, Q6 = rno . i; = m3. (5.2.9)

Thus, the generalised forces corresponding to the Euler's angles 'ljJ, {}, t.p are
the projections of the principal moment on axis Oz, the nodal axis and axis
Oz' which are the principal moments about these axes.
We now express the scalar products in eq. (5) in terms of the projections
of the vectors on axes Ox' y' z' fixed in the body

b"W = VI (b'rO)l + V2 (b'r O)2 + V3 (b'rO)3 + m?(h + m~(h + m~83.


(5.2.10)

Here (b'rO)l , ... ,83 are variations of the quasi-coordinates, corresponding to


the quasi-velocities introduced by relationships (1.5.2) and (1.5.3). Adopt-
ing the numbering used in Sec. 2.10, i.e. assuming that

b'11"1 = 81 , b'11"2 = 82 ,
(5.2.11)
b'11"4 = (b'rO)l , b'11"5 = (b'rO)2'

we obtain

PI = ml, P2 = m2, P3 = m3, P4 = VI, P5 = V2, P6 = V3.


(5.2.12)
5.2 Elementary work of forces acting on a rigid body 207

Thus, the generalised forces corresponding to the quasi-coordinates (11)


are the principal moments of the system of forces about the axes fixed in
the body and the projections of the principal force on these axes. While
writing eqs. (8) and (12) we used the notation ml instead of m~, since the
projection of the moment rn D on a certain axis, that is a moment about
this axis, does not depend on a particular point on this axis.
Provided that the projections of the angular velocity vector on axes Oxyz
fixed in the space are taken as the quasi-velocities, the corresponding gen-
eralised forces are the principal moments m!, m2, m3 about these axes.
Let us relate these quantities with the moments m3, mN, m3 which are
the generalised forces corresponding to Euler's angles. We have

(5.2.13)

but, of course,

rn D i=- m3 i3 + mNll + m3i~,


since the directions h, ll, i~ do not form an orthogonal trihedron. The cor-
rect representation is as follows

rn D = mNll+~ [(m3 - m3 cos 19) i~ + (m3 - m3 cos 19) i 3], (5.2.14)


sm 19
as can be proved easily by means of eq. (9). Indeed, scalar multiplication of
eq. (14) by ll, i~, i3 yields the above formulae. A more detailed explanation
of the difference between the vector components in axes of a non-orthogonal
coordinate system and its projections on the axes is given in Sec. B.2.
Using eqs. (2.3.1)-(2.3.3) we obtain, from eq. (14),

(5.2.15)

and from (13) and (9) the inverse relationships

mN =mlcos'ljJ+m2sin'ljJ, }
m3 = ml sin 19 sin 'ljJ - m2 sin 19 cos 'ljJ + m3 cos 19, (5.2.16)
m3 = m3·
By analogy we have

(5.2.17)
208 5. Work and potential energy

and the inverse relationships

mN = m1 COS<p - m2 sin<p, }
m3 = m3, (5.2.18)
m3 = (m1 sin <p + m2 cos <p) sin iJ + m3 cos iJ.

5.3 Potential energy


The concept of elementary work due to the virtual displacement of the
points of the system was introduced above. By analogy we define the ele-
mentary work done by the actual displacements as
N
d'W = LFi · dri. (5.3.1)
i=l

Using this quantity we define the work of the forces during a finite dis-
placement of each point in the system as the following integral
(2) (2) N

W 12 = jd'W= j~Fi.dri' (5.3.2)


(1) (1) .=1

whose limits are given by the values of the coordinates of the points of the
system at the initial (1) and final (2) positions. If the motion is prescribed
this integral can be evaluated since the integrand becomes a given function
of time, while the limits of the integral are determined in terms of the time
instants t1 and t2, corresponding to the initial and final positions of the
system.
Let us consider the case where the forces depend only on the positions of
the points referred to an inertial coordinate system. These forces are termed
potential provided that there exists a single-valued and twice-differentiable
function with respect to all of its arguments

(5.3.3)
such that the projections of the force acting on the point Mi(Xi, Yi, Zi) are
equal to the negative partial derivatives of n with respect to the corre-
sponding coordinate, i.e.
an an an
Fix = --a
Xi
' F iy = --a'
Yi
F iz = --a'
Zi
(5.3.4)

It is assumed that function n depends upon only the point coordinates


and that time does not appear explicitly in expression (3). It follows from
eq. (4) that this function is determined up to an additive constant and is
referred to as the potential energy.
5.3 Potential energy 209

The vector whose projections on the coordinate axes are equal to the
partial derivatives of a function tp with respect to the coordinates Xi, Yi, Zi
of a point Mi is called the gradient of the function tp at this point and is
denoted by gradi tp. For this reason,

(5.3.5)

In accordance with eqs. (1) and (4), the elementary work of the potential
forces due to actual displacements is given by

d'W = - L grad II . dri = - L


N
i
N (all -dXi
all
+ -dYi + -dzi all) = -dIl,
i=l i=l aXi aYi aZi
(5.3.6)

where dIl is the total differential of the potential energy.


The work of the potential energy in a finite displacement is, due to eqs.
(2) and (6), given by

J J
(2) (2)

W 12 = d'W = - dIl = III - Il 2 . (5.3.7)


(1) (1)

It follows from this expression that the potential energy is equal to the
work which would be done by the potential forces in a finite path from the
considered position of the system to a position where the potential energy
is assumed to be equal to zero.
Let us consider the case of stationary constraints. Due to eq. (1.2.11)
time does not appear in expressions for the Cartesian coordinates in terms
of the generalised ones and thus the potential energy becomes a function
of the generalised coordinates only

(5.3.8)

According to eq. (6) the elementary work of the potential forces due to
a virtual displacement is

(5.3.9)

Comparing this with eq. (1.4) and taking into account that the variations
of the generalised coordinates are independent, we arrive at the following
expression for the generalised forces in terms of the potential energy

Qs = - all (s = 1, . .. ,n) . (5.3.10)


aqs
210 5. Work and potential energy

Assume that the generalised forces Ql, ... , Qn are known and that they
depend only on the generalised coordinates Ql, ... , qn' As follows from the
latter equation, an indication of a potential force is that it meets the fol-
lowing condition

(k,s=l, ... ,n). (5.3.11)

If this condition is not met, then the forces are not potential. If this
condition is satisfied, then the elementary work d'W is a total differential
of a function of the coordinates and the forces are potential forces provided
that this function is single-valued. For example, let the projections of force
F on axes Oxy be
Y Y
Fx = -k 2 2' Fy = k 2 2'
X +y X +y
It is easy to prove that condition (11) is met and that the elementary work
is given by

d'W = k (- 2 Y 2dx + 2 y 2dY) = kdarctan J!....


x +y x +y x
Since function arctan(yjx) is not defined at the origin 0, force F is not
potential in the region which includes point O.
Introducing quasi-coordinates we have, due to eq. (1.8),
n n all
P s LbrsQr = - Lbrs8 (5.3.12)
r=l r=l qr
or recalling notation (1.5.17)

P s -__ nall (S = I, ... , n ) . (5.3.13)


U7rs

Sometimes the generalised forces can be represented by a formal equality

Qs __
-
all (
s=1 , ... ,n,
)
aqs
which is analogous to eq. (10). In the latter equation the function II can
depend not only on the generalised coordinates but also on the time

(5.3.14)

This is particularly the case for nonstationary constraints when expres-


sions (1.2.9) for Cartesian coordinates in terms of the generalised coordi-
nates are substituted into eq. (3). Another example is a force depending
only on time, that is

Fx=f(t),
5.3 Potential energy 211

Then

II (x, t) = - f (t) . x.
Function (14) is referred to as the generalised potential energy.
The elementary work of the forces due to actual displacement is expressed
in terms of the potential energy as follows

(5.3.15)

and therefore is not a total differential of function (15). Equation (7) is also
not fulfilled.
Later it will be made clear what is meant by potential energy and gener-
alised potential energy. With this in mind we omit the word" generalised"
for the purpose of notational convenience.
A simple example of potential forces is a force of constant value and con-
stant direction. For example, the gravitational force in the neighbourhood
of the earth's surface. The potential energy of the system in the gravita-
tional field is given by the expression
N

II =9 L miZi = Mgzc, (5.3.16)


i=l

where Zc denotes the coordinate of the centre of gravity along the upward
vertical, with the origin on the earth's surface. Indeed, expression (16)
is equal to the work which would be done by the gravity forces due to
displacements of the system particles from the actual position to the earth's
surface.
Expression (16) can also be represented in the form
N
II = -Mg· rc = -g. L miri, (5.3.17)
i=l

where g is the vector of free fall acceleration directed along the downward
vertical and r c denotes the position vector of the centre of inertia of the
system. In a more general case the potential energy of a system of forces F i
of constant value and constant direction acting on points Mi of the system
is determined by
N
II =- L Fi . rio (5.3.18)
i=l

As an example let us consider a system of two rods 0 1 0 2 and 0 2 0 3


with joints at points 0 1 and O 2 , the free end 0 3 of the second rod being
212 5. Work and potential energy

FIGURE 5.1.

subject to a constant force T , see Fig. 5.1. The position of the system is
determined by two generalised coordinates: the angles 'PI and 'P2 between
the rods and the axis OIX. The angle between the direction of force T and
axis OIX is denoted by {3.
The elementary work of force T is

8'W Tx8x03 + T y8Y0 3


T [cos {3<5 (h cos 'PI + h cos 'P2) + sin (38 (h sin 'PI + l2 sin 'P2)]
and therefore

i.e. the generalised forces are equal to

(5.3.19)

These will be the generalised forces under condition (11) , i.e.

h cos ({3 - 'PI) ~{3 = h cos ({3 - 'P2) ~{3 . (5.3.20)


u'P2 u'PI
In particular, the latter condition is satisfied when {3 = const, i.e. the force is
in a constant direction in addition to the above requirement of the constant
value. This force is potential force and the potential energy is given by

II = -T [h cos ({3 - 'PI) + l2 cos ({3 - 'P2)]· (5.3.21 )

In the case of a force T having a constant value and a constant angle a


between its direction and the rod 0 2 0 3 , we have {3 = a + 'P2 and then

QI = Th sin (a + 'P2 - 'PI)' Q2 = Tl2 sina.


5.3 Potential energy 213

This follower force is an example of a so-called positional force which is


not a potential force.
Returning to formula (18) let us consider two positions of a rigid body
having a fixed point 0, namely an initial position and a final position which
is obtained from the initial one by a finite rotation e. The forces F~, ... , F~
acting on the body points with the position vectors r~, ... , r~ hold their
values and directions in space under the above rotation. An expression for
the potential energy of this system of forces is required.
Using Rodrigues formula, the initial position vectors r~ becomes, after
rotation

and byeq. (17)


n

The term ITo, denoting the potential energy in the initial position, can be
cancelled out, whilst the moment of force F~ in the initial position about
the pole 0 is given by

Thus, the result can be cast in the form

IT = - ()
1 + 1()2
. ° + -21 1 +11()2 e· Q ° . e.
rna (5.3.23)
4 4

Here rn~ denotes the principal moment of the system of forces about the
pole 0 in the initial position and QO denotes the following tensor
n
QO = L (EF~ . r~ - F~r~) . (5.3.24)
k=1

Expressing the vector e in terms of the Rodrigues-Hamilton parameters


by means of formulae (3.2.8) and (3.2.10), we can rewrite formula (23) as
follows
3 3 3

IT = - 2A o L Asrn~ . i~ + 2 L L QskAsAk. (5.3.25)


s=1 s=1k=1
214 5. Work and potential energy

Here i~ stands for the unit vectors of the trihedron bound to the body in
the initial position and Qsk denotes the components of tensor QO with
respect to these axes.
Let, for example,

F o = F'o
11'

Then,

°
rno = Zo F'1o
2,

and inserting the Rodrigues parameters (3.6.6) in terms of Euler's angles


into eq. (25) we obtain

II -2 (AOA2 + A1A3) Fzo + 2 (A~ + A~) Fxo


- Fzo sin {) sin 'ljJ + Fxo (1 - cos'ljJ cos cp + sin'¢ sin cp cos '!9) .

Of course, the same result can be obtained by means of the formula

II = -F (x - xo),

where xo and x denote the initial and final coordinates of the point on
which the force acts, respectively. This coordinate can easily be found by
means of Table 2 of the direction cosines from Chapter 3

x = xo (coscpcos'ljJ - sincpsin'ljJcos'!9) + zosin'!9sin'ljJ.

5.4 Forces that depend linearly on the coordinates


We consider now an example of positional force F whose projections on
axes of the Cartesian coordinate system OX1X2X3 are prescribed by linear
functions of the coordinates

(5.4.1)

where Psk are constant.


Introducing into consideration a tensor P and the position vector r, we
can write the latter equation in the form

F=P·r, (5.4.2)

see (A.4.lO). Let pI denotes the transpose of tensor P, i.e. such a tensor
that Plk = Pki. The relationship

p = ~ (P + pI) + ~ (P _ pI) (5.4.3)


5.4 Forces that depend linearly on the coordinates 215

defines splitting of tensor P into the symmetric part

(5.4.4)

having the components

(5.4.5)

and the skew-symmetric part

(5.4.6)

where

The expression for the force F is written in the form

F = S ·r+Q·r, (5.4.7)

where by virtue of eqs. (1) and (6) the second term, i.e. the vector Q . r,
has the following projections on the coordinate axes

We introduce into consideration a vector 0 whose projections on axes


OXIX2X3 are respectively 0 1 , O2 , 0 3 , then

F = S· r+O x r. (5.4.8)

This notation is justified by the fact that under the transformation of the
coordinate system values Os are transformed like the vector projections, see
Sec. 2.12 for detail. Due to relationships (5), the first term is the gradient
of the quadratic form

(5.4.9)

i.e. it is the potential part of the force F

F = - grad II + 0 x r. (5.4.10)

The second term is a non-potential force called the circulatory force or the
force of radial correction, the latter concept being used in the theory of
216 5. Work and potential energy

FIGURE 5.2.

gyroscopes. It is directed perpendicular to the position vector of the point


at which the force acts. In a simple case its projections are given by

Fx = -ky, Fy = kx. (5.4.11)

As follows from the hydrodynamic theory of lubrication, the principal vec-


tor of reaction of the oil film on the rotating shaft is an example of a
circulatory force. By Sommerfeld's theory

F = 127fp,R3 W x r. (5.4.12)
E2 (2+>.2) ~
Here R' and R denote the radius of the bearing and the shaft, respectively,
€ = R' - R , p, is a factor of the lubrication viscosity, A = E/r, and w the
angular velocity vector of the shaft. In contrast to (11) the proportionality
factor is dependent on r. Figure 5.2 shows the force F as applied at the
shaft centre 0' . However it is well to bear in mind that the reaction of
the oil film on the shaft is statically equivalent to the principal vector F
and the principal moment m O ', the latter being proportional, and in the
opposite direction to, the vector w. Thus the point at which the resultant
of the oil film reaction forces is applied does not coincide with 0'.

5.5 Potential energy due to the force of gravity


The law of gravitation states that particle M of mass m is attracted by a
particle of mass mo (the attraction centre) with the gravity force

F = _fmmo~ (5.5.1)
r2 r'
5.5 Potential energy due to the force of gravity 217

where r denotes the distance between the bodies and r the position vector of
point M with respect to the centre of attraction. The universal gravitational
constant f is equal to

(5.5.2)
The elementary work of the force F due to the virtual displacement 8r
is
, mmo
8 W = - f-2-r . 8r.
r
Since
1 1 2
r . 8r =-8 (r . r) = -8r = r8r
22'
we have
8'W = - fmm o8r = 8 (fmm o ).
r2 r
It follows from eq. (3.9) that the potential energy of the force of attraction
of two particles is described by the expression

(5.5.3)

We proceed now to calculate the potential energy of the particle M


attracted by a rigid body of finite dimensions which is far removed from
M. The origin of the coordinate axes is placed at the centre of inertia G of
the body and axes Gx, Gy, Gz are assumed to coincide with the principal
central axes of inertia. Then, by virtue of eq. (3)
N
IT = - fm '"'" m·' ,
L...J - (5.5.4)
i=l ri

where mi denotes the mass of particle Mi in the attracting body, and ri the
distance from the attracted particle M. The position of the particle Mi in
the b0:J is given by the position vector Pi = GM,
see Fig. 5.3. Denoting
r = G we obtain

r~, = r2 + p~, - 2r . P',


and

(5.5.5)

Furthermore

r . Pi = XXi + YYi + ZZi,


218 5. Work and potential energy

FIGURE 5.3.

where x, y, Z and Xi, Yi, Zi denote the coordinates of points M and M i ,


respectively. Expanding (5) as a second order polynomial in p/r, this ratio
being assumed to be small, we arrive at the following equality

1 1 ( r· 1 PT
ri
-
r
1+ --
r2
Pi
- --
2 r2
+ -32 (r .r Pi)2
4
+ ... )

1 XXi + YYi + ZZi 1 x; + YT + z; 3 (XXi + YYi + ZZi)2


-:;. + r3 - "2 r3 + "2 r5 + ...
While inserting this expression into (4) it is necessary to consider that the
coordinate origin G is taken at the centre of inertia which implies that
N N N
L miXi = L miYi = L miZi = 0,
i=l i=l i=l

whereas the coordinate axes coincide with the principal axes of inertia such
that
N N N
L miXiYi = L miYiZi = L miZiXi = o.
i=l i=l i=l

Hence we obtain

II =

Here mo denotes the total mass of the body. Due to eq. (4.3.21)
5.5 Potential energy due to the force of gravity 219

N 1
L mix; = "2 (8 2 +8 3 - 8I)
i=l

and so on, 8 1, 8 2 , 8 3 being the principal moments of inertia, we obtain

II =

The force F* acting on the attracted particle M is determined by eq.


(3.6). Due to the law of action and reaction, the particle M exerts a force of
attraction F upon the body. The absolute value of F and F* coincide, and
force F acts along the straight line joining the centre of inertia of the body
and point M. Thus, a body of finite dimensions in the field of an attracting
centre which is a particle is subjected to attraction forces whose principal
force and principal moment about the centre of inertia of the body are
equal to

F = -F* = gradll, mG = rx gradll, (5.5.7)

respectively. Thus the projections of the principal vector on axes Gxyz are

(5.5.8)

where
all
ar

The principal moments about these axes are given by

(5.5.10)

When considering the rigid body motion we should express the forces and
moments acting upon the body in terms of the parameters determining the
220 5. Work and potential energy

a)

FIGURE 5.4.

position of the body whereas the resulting formulae contain the coordinates
of the attraction centre M referred to the axes fixed in the body. For this
reason, the fixed axes M~'r/( with the origin in the attracting centre are
introduced, Fig. 5.4a. The position of the centre of inertia G with respect
to this system is determined by the position vector MG = -r or by the
coordinates ~,'r/, (. The directions of axes Gxyz are described by Euler's
angles which determine the direction cosines Qik of these axes with respect
to the fixed axes.
Projecting vector r onto axes Gxyz we have

x = - (Qll~ + Q12'r/ + Q13() , }


y = - (Q21~ + Q22'r/ + Q23() , (5.5.11)
Z =- (Q31~ + Q32'r/ + Q33().
These relationships should be substituted into expression (6) for the po-
tential energy or directly into formulae (8) and (10). This would yield very
bulky expressions for the generalised forces Ps in terms of the generalised
coordinates. More transparent results are obtained when the central el-
lipsoid of inertia of the body is an ellipsoid of revolution. Let Gz be the
axis of revolution and the positive direction be taken such that the angle
e between axis Gz and vector eM = r is positive, see Fig. 5.4b.
As 8 1 = 8 2 we have

m z = O. (5.5.12)

Due to eq. (7) the vector m G , which is perpendicular to r, is now per-


pendicular to axis z, i.e. the unit vector i~. This can be cast in the form

mG = 3fm (8
--5- - 8 1 ) Z 13
./ x r. (5.5.13)
3
r
5.6 The shape of the Earth 221

As z > 0 and provided that 8 3 < 8 1 (which implies that the ellipsoid
is extended along the axis of rotation) the moment m G tends to bring the
axis Gz into coincidence with the direction at the attraction centre M.
Since m z = 0, the force F acts in the plane passing through axis Gz and
point M. Projections of F on the axes, due to eqs. (8) and (9), are given by

Fx = fm [mo + ~4 (8 3 - 8d (1- 5cos2 ())] ::,


r2 2r r

F =
y
fm [mo
r2
+~
2r 4
(8 3 - 8 1) (1 - 5cos2 ())] 1J..,
r
(5.5.14)

Fz = fm [;20+ 2~4 (83- 8 1)(3 - 5 cos 2()) ] cos ().


We now enter the orthogonal trihedron of the unit vectors n, n' ,i; "half-
fixed" in the body. The vector n is orthogonal to the plane of rand i;,
while n' is directed along the component of r in the plane Gxy as shown
in Fig. 5.4b. Then

(5.5.15)

(5.5.16)

In the case depicted in Fig. 5.4b, i.e. 8 3 < 8 1, the moment mn < O.

5.6 The shape of the Earth


The question of the shape of the earth's surface is relevant to the theory of
gyroscopic devices, problems concerning the motion of satellites, and other
spheres of dynamics. On the other hand, this question provides us with an
excellent example of the application of the concept of potential energy. The
contents of this section is based on [40j.
The surface of a normal spheroid approximates the earth surface and is
the surface of an equal level of the gravity force which is a sum of the earth
gravitation and the centripetal force due to the earth rotation.
A normal spheroid differs slightly from a sphere, and the forthcoming
analysis is limited by the first order of small parameters characterising this
difference.
Let the spherical coordinates r, fJ, >. of a point denote the distance from
the centre of the earth, the angular distance between the north direction
222 5. Work and potential energy

and the meridian (the complementary angle to the latitude) and longitude,
respectively. We assume that the normal spheroid is a body of revolution
with the rotation axis Oz and denote its equatorial and polar moments of
inertia as A = 8 1 = 8 2 and C = 8 3 , respectively. In accordance with eq.
(5.6) we have the following expression for the specific potential energy (i.e.
m = 1) of the gravity force

-fM [ -1 +- A (2
C -- x + y 2- 2z 2)]
r 2Mr 5
1
- f M [ ;: + C2Mr
- A
3 (sin 2 {} - 2cos 2 {})
]
. (5.6.1)

Here M denotes the mass of the earth. As it will become clear later, cf.
eq. (9.2.6), while considering a motion or an equilibrium with respect to
axes rotating along with the earth one should consider the potential field
of the centripetal forces. The specific potential energy of the centripetal
forces due to the rotation of the earth is

(5.6.2)

with U being the angular velocity of the earth. Summing up eqs. (1) and
(2) and introducing an average radius of the earth Ro we arrive at the
equality

II f M {Ro C - A (Ro) 3 U 2R~ ( r ) 2


- Ro --;:- + 2M R6 --;:- + 2f M Ro

[ 3(C-A)
2MR6
(Ro)3
r +
U2R~(.!..-)2]
2fM Ro cos
2{)}.
The following notation

3 (C - A) U2R~
2MR6 + 2fM = 0:, (5.6.3)

is adopted in the theory of the shape of the earth. It enables one to write
the above expression for the potential energy in the form

II _fM {RO +~ (0: _ m) (Ro)3 + m(.!..-)2

r (~o r]
Ro r 3 2 r 2Ro

[(0:- ;) (~ +; cos 2 {} } . (5.6.4)


5.6 The shape of the Earth 223

This is equivalent to the expression

IT =

in which P2 (cos '!9) denotes the second polynomial of Legendre


1
P2 (cos '!9) = "2 (3 cos 2 '!9 - 1) . (5.6.6)

In what follows, the products and the sqnares of small values m and a
are omitted since it is consistent with discarding the higher order terms in
eq. (5.6) for the potential energy of the gravitational force.
The value
fM
-2- =90 (5.6.7)
Ro
represents the acceleration of an "average gravity force" which is the at-
traction force of the non-rotating earth at radius Ro. With this in view the
value
U2RO
m=-- (5.6.8)
90
is the ratio of the centripetal force on equator to the average gravity force.
The above value Ro is now defined as the radius ~o of such a sphere
that its specific potential energy ITo is equal to the average value of the
potential energy on the surface of this sphere

where

do = R5 sin '!9d'!9d)"
denotes the element of the sphere surface ~o. Hence, due to eq. (5)
224 5. Work and potential energy

since by virtue of the Legendre polynomials

J
7r

P2 (cos'!9) sin '!9dTJ = O.


o
Instead of II we will consider its deviation from IIo

II - IIo -~ {; -1 + ~ [G~f -1]-


~ [(a - ;) (~o) 3 +; (~rl P 2 (COS'!9)} (5.6.9)

The normal spheroid of Clairaut is defined as a body on whose surface


the potential energy is equal to the potential energy of the sphere ~o which
is an "average spherical earth". This definition introduces the distance R*
between the centre and the surface of the normal spheroid as follows

(II - IIo)r=R* = O. (5.6.10)

This equation determines R* as a function of angle TJ. The solution is sought


in the form

R* = Ro [1 +'Y(TJ)].

The quantity 'Y (TJ) describing deviation of the normal spheroid from the
average spherical earth has the order of smallness of a and m. Retaining
only terms of first order we have, due to eq. (10),

and the equation for the surface of the normal spheroid takes the form

(5.6.11)

At the equator and the poles, i.e. at TJ = 7r /2 and TJ = 0

(R*)1'J=7r/2 = a = Ro (1 + ~), (R*)1'J=o = C = Ro (1 - ~a). (5.6.12)

The values a and c are referred to as the equatorial and the polar radii of
the spheroid and
a-c a-c
a=--~--
Ro a
is called its oblateness.
5.6 The shape of the Earth 225

The surfaces of the normal spheroid and the average spherical earth
intersect at the parallel circles for which P2(1')) = 0, i.e. at

cos 1') = ±/f, 1')1 = 35°16', 1')2 = 144°44'.

The equation for the normal spheroid can also be written in the form

R* = Ro (1 + ~ - a cos 2 1')) = a(1 - a cos 2 1')) . (5.6.13)

Let 1')* denote the angle between the normal to the surface of the normal
spheroid and its rotation axes Oz. If p and z denote the cylindrical coor-
dinates of the point, then on the surface of the normal spheroid they are
given by
p* = a (1 - acos 2 1')) sin 1'), z* = a (1 - acos 2 1')) cos 1').

Then
dz* 1 - 3acos 2 1')
tan1')* = - - = 2 21') tan 1') ~ (1- 2a)tan1')
dp* 1 + a - 3a cos
or
1')* -1')
tan 1')* - tan 1') ~ -----z-::a = -2a tan 1'), E = 1') - 1')* = a sin 21'). (5.6.14)
cos u
We proceed now to calculating the force in the potential field in question.
With the help of eq. (3.6) we find the force projections on the axes of the
spherical coordinate system

Fr = - all
or
= -go [(Ro)2
r
- ~m~-
3R o
R4 2 r ]
(2a - m) -f:P2 (cos 1')) + -m-P2 (cos 1')) ,
r 3 Ro (5.6.15)

F{}=--=go all [(a -m)


- (Ro)
- m
4 +-- r 1sm21'),
.
ro1') 2 r 2 Ro
FA =0,
where go is given by eq. (7). At the spheroid's points we have

F; = -go [1 + ~ - ~m - (a - %m ) cos 2 1')], } (5.6.16)


FJ = goa sin 21').
The components of the gravity force along the normal n to the spheroid
and the tangent T to the meridian are as follows

F~ F; COSE - FJ sinE ~ F;,


F; F; sinE + FJ COSE = -goa sin 21') + FJ = 0.
226 5. Work and potential energy

The latter result serves to validate the calculation validity because the
vector grad II is normal to the surfaces II = canst and the spheroid is
bounded by such a surface. The expression for F; can also be written in
the form

F~ = -ge (1 + (3cos2 'l9), (5.6.17)

where

(5.6.18)

stands for acceleration of the gravity force at the equator and the Clairaut
constant is
5
(3 = -m
2 - a. (5.6.19)

We also construct an equation for the normal spheroid by using cylin-


drical coordinates. We have due to eq. (13)
2
2 2 2
R* =p* +z* =a 2 (1-acos 'l9) 2 2 2
~a (1-2acos 'l9)~a -2az
2 *2

or
*2 *2
P
-2
a
+ -Za2 (1 + 2a) = l.

By virtue of eq. (12)

and thus
*2 *2
P Z
-2 +-2 =l. (5.6.20)
a c
In the framework of the adopted accuracy the surface of the normal
spheroid of Clairaut and the normal ellipsoid of Clairaut coincide.
To conclude we present some numerical data. The constants a and a,
defining the size and the shape of the normal spheroid of Clairaut, have
the following numerical values
1
a = 6378.4 km, a ~ 296.3 ~ 0.003375. (5.6.21)

The constants m and (3 are equal to


1
m = 288.4 = 0.00347, (3 = 0.00529. (5.6.22)
5.7 Elastic forces 227

By means of (3) we obtain


3C-A
2 M R5 = 0.00164. (5.6.23)

The equinox anticipation theory yields


C-A
-C = 0.003275. (5.6.24)

Comparing the latter two equations we obtain


C
MR2 = 0.334 (5.6.25)

instead of 0.4 for a homogeneous sphere. This indicates that the average
density of the earth increases towards its centre.
The difference between geographic 7r /2 - {)* and the geocentric 7r /2 - {)
longitudes is given by eq. (14). Its maximum values is achieved at longitude
45° and is equal to a. A more accurate analysis yields E"max = 1/282.

5.7 Elastic forces


Since analytical mechanics is restricted to the analysis of the motion of
systems with a finite number of degrees of freedom, one should consider
elastic bodies as a massless source of effects which are the elastic reactions
to points of the material system attached to the elastic body. In Chapter
12 we show some examples of approximate methods based upon replacing
the solid by models whose configuration can be sufficiently well described
by a finite number of parameters.
The property on an ideally elastic body is that its reactions caused by
changes in the form and size depend only on the quantities determining
the position of the system points and do not depend upon their time-rates
and the time-history of the deformation. For this reason, in the sequel we
assume that the generalised forces of the elastic reactions are functions only
of the generalised coordinates of the system
(5.7.1)
Elastic forces are potential forces. This physical assumption expresses the
property of elastic bodies to accumulate potential energy under a mono-
tonic loading and to give it back without any loss when the body, under a
monotonic unloading, returns to the natural state, in which its form and
size are restored.
Hence, the coordinate functions in eq. (1) are assumed to be represented
in the form
Q __ orr 1 (5.7.2)
s - oqs s = , ... , n,
228 5. Work and potential energy

where II (ql, q2, ... ,qn) is the potential energy of the elastic bodies com-
prising the system. The sequence of eq. (2) is the above relationships (3.12)

(5.7.3)

which the generalised potential forces should obey to.


rt is assumed in what follows that the generalised coordinates are mea-
sured from that configuration of the system in which the elastic bodies are
in the natural state. In other words, the elastic reactions vanish when the
generalised coordinates are equal to zero

Qs (0, ... ,0) = O. (5.7.4)

The main physical law of the mathematical theory of elasticity is the


generalised Hooke '8 law expressing the linear relationships between the
quantities describing the stress state, i.e. normal and shear stresses and the
quantities describing the deformations, i.e. elongations and shear strains.
This property of the ideally elastic body is observed for the majority of
materials under small deformations.
Can we assert also that Hooke's law determines a linear dependence
between the displacements of the points of the elastic body and, as a con-
sequence, a linear dependence between the generalised forces of the elastic
reactions and the generalised coordinates? Numerous simple examples give
evidence that this is not always the case. For example, the reaction of an
elastic medium to a rigid sphere is not linearly proportional to the dis-
placement of the sphere. The end of a vertical rod subjected to an axial
compressive force exceeding Euler's critical value moves downwards as well
as sideways. This displacement is not small though the strains are small
and Hooke's law is not violated.
Despite these examples a great number of problems of the theory of
elasticity, based on Hooke's law, indicate linear relationships between the
forces acting on the body and the displacements of its points. This justifies
adoption of a modified generalised Hooke's law stating that the elastic force
depends linearly on the coordinates
n
Qs = - L Cskqk = - (Cslql + ... + csnqn) (8 = 1, ... ,n) , (5.7.5)
k=l

where Csk are constant coefficients of the stiffness matrix

C= (5.7.6)
Cnl Cn 2 Cnn
5.7 Elastic forces 229

Since the elastic forces are potential forces, this matrix is symmetric.
Indeed, as follows from eqs. (3) and (5)

Csk = Cks (k, s = 1, ... , n) . (5.7.7)


It is clear that eqs. (5) are applicable only for sufficiently small values
of the generalised coordinates measured from the natural configuration of
the system. It corresponds to the prerequisite for small strains and the
assumption that the elastic system does not allow considerable changes in
the form and size.
Due to eqs. (2) and (5) variation of the potential energy of the elastic
forces is equal to

(5.7.8)

An expression for the potential energy is easy to recover with the help of
the total variation (8). However this calculation would be unnecessary since,
due to eq. (8), the potential energy is a quadratic form of the generalised
coordinates (the constant term can be left out). The bilinear expression for
this form in terms of the generalised coordinates and generalised forces is
easily constructed by using the theorem on homogeneous functions (4.1.12)

(5.7.9)

then, by virtue of eq. (5), we obtain


1 1
L L cskqkqs = 2"q' cq.
n n
II = 2" (5.7.10)
s=l k=l

Here q and q' are a column matrix and row matrix of the generalised
coordinates, respectively, and similar notation for Q and Q' is used for the
generalised forces.
A typical property of an elastic body is that the work of the forces
of elastic reactions is positive when recovering the natural configuration,
which implies positiveness of the potential energy at any configuration other
than the natural one. For this reason the quadratic form (10) is a positive
semi-definite. One can assert that this form is positive definite function
of the generalised coordinates only under certain conditions, see Sec. 1.3.
First of all it is necessary to agree that ql, . .. , qn in eq. (10) and the other
equations of this section designate only those parameters which are present
in these equations but not all the independent parameters describing the
system configuration. These parameters should be taken so that they should
all vanish in the natural configuration of the elastic bodies comprising the
230 5. Work and potential energy

FIGURE 5.5.

system. An example illustrating this condition is depicted in Fig. 5.5. A


rigid plain rigid body S is attached to an elastic rod by means of a joint
at point O. The other end of the rod is clamped. The position of the
body in the plane perpendicular to the joint axis is described by means of
three parameters, for example the projections Ua = ql and Va = q2 of the
displacement of point 0 on axes Ox and Oy, respectively and the rotation
angle <p = q3. In this case the potential energy II will be a positive definite
quadratic form of ql and q2 , with q3 not appearing in this expression.
However the projections qi = U M and q2 = V M of the displacement of point
M on axes M x* and M y*, respectively, and the rotation angle <p = q3 can
also be taken as the generalised coordinates. Then for small values of <p we
have

and II ·becomes a positive semi-definite function of three variables qi, q2 ' q3·
This form is not positive definite as it is equal to zero at qi = -bq3, q2 =
-aq3'
From here on it is assumed that II is a positive definite quadratic form of
the generalised coordinates. The elements Csk of the stiffness matrix satisfy
the Sylvester inequalities (A.3.25) . One of them, namely the positiveness
of the determinant Ici of the quadratic matrix, ensures the existence of the
inverse matrix

(5.7.11)
5.8 Calculation of the potential energy for rod structures 231

referred to as the influence matrix. It enables us to inverse eq. (5)


n
qs = - L askQk = - (a s 1Q1 + ... + asnQn) (s = 1, ... ,n). (5.7.12)
k=l

The inverse of a symmetric matrix is also symmetric, i.e.

ask = aks (k, s = 1, ... ,n). (5.7.13)

This equation expresses Maxwell's reciprocal theorem.


Substituting qs in terms of the generalised forces into a bilinear expres-
sion for the potential energy (9) yields another representation for rr in
terms of the generalised forces

(5.7.14)

By virtue of eqs. (12) and (14) we arrive at Castigliano's theorem

orr
- aQ = - L askQk = qs
n
(s = 1, ... ,n). (5.7.15)
s k=l

5.8 Calculation of the potential energy for rod


structures
Calculation of the influence and stiffness matrices is performed using the
theory of elasticity and structural mechanics. A few examples are elabo-
rated in this and the following sections.

5.8.1 A statically determinate system


A rigid plate S attached to two trusses at points 0 1 and O 2 by means of
two joints is considered. Each truss is a statically determinate system of the
rods hinged at their ends. In the equilibrium position the rods are stressed
since firstly they react the load applied to the plate (e.g. the plate weight)
and secondly the natural lengths of the rods may differ from their actual
length in the assembled structure. We ignore the fact that the structure
can be initially prestressed and take the equilibrium position depicted in
Fig. 5.6 as the natural configuration of the elastic system.
It is assumed that the plate moves in its plane and the displacement
from the equilibrium is small. The rotation angle cp and the coordinates
XG, YG of the pole G are taken as the generalised coordinates. The latter
are counted with respect to a fixed coordinate system Goxy whose origin
232 5. Work and potential energy

-0 o

FIGURE 5.6.

coincides with the initial position of the pole, see Fig. 5.7. The projections
of the pole displacements on axes Gx'y' fixed in the plate

Xc Xc cos <p + Yc sin <p ~ Xc + Yc<P ,


Yc -Xc sin <p + Yc cos <p ~ -xc<p + Yc,

differ from Xc and Yc only in the second order terms. For this reason, we
identify these quantities. Up to the same order of accuracy the projections
of displacements of hinges 0 1 and O 2 on axes Goxy are equal to

X0 1 = Xc + b<p, YOl = Yc - a<p, x0 2 = Xc - b<p, Y0 2 = Yc + a<p.


(5.8.1)
We proceed now to construct expressions for the potential energies of
the elastic reactions of the left and right trusses in terms of X0 1 ,Yo 1 and
X0 2, Y02' respectively.
We mentally load the truss shown in Fig. 5.8 using a unit force X = 1
along the positive axis x. By using methods from structural mechanics we
determine the forces in the rods. Let Ak denote the force in the k - th rod.
By analogy we find the forces Ji.k due to the unit load Y = 1 shown in Fig.
5.8. Given Ak and Ji.k , we find the forces Sk in the rods of the truss subject
to arbitrary forces X and Y at hinge 0
(5.8.2)
y

~ ~'
~---
6'0'--------;&

FIGURE 5.7.
5.8 Calculation of the potential energy for rod structures 233

Y-I

0, X-'

FIGURE 5.8.

as well as the potential energy of the truss as the sum of the potential
energies of all m rods

Here Fk and Ek denote the cross-sectional area and Young's modulus of


the k - th rod, respectively. We denote the influence coefficients determined
by the geometry and material of the truss as

m A% m 2

au = 2:: E F
k=l k k
' a22 =
~
~
k=l
ILk
E F
k k
. (5.8.4)

Since the external forces X and Y can be understood as the generalised


forces -Ql and -Q2 corresponding to the generalised coordinates Xo and
Yo (horizontal and vertical displacements of the hinge 0) , we obtain

(5.8.5)

Because

the expression for II in terms of the generalised coordinates Xo and Yo takes


the form

II 1 ( CuXo2 + 2C12XOYO + C22Yo2)


=2 , (5.8.6)

where the coefficients of the stiffness matrix are


234 5. Work and potential energy

FIGURE 5.9.

Returning to the example of Fig. 5.6 we have

II = "21(, 2 2'
Cll XO + C12 XO
l l YO l
, YO2 ) +
+ C22 l

"21(" 22 + 2"
Cll X0 "2)
C12 X02 Y02 + C22Y0 2 (5.8.7)

where <k and <~ are coefficients of the stiffness matrices of the left and
right trusses, respectively. Expressing by means of eq. (1) XO l , .•• , Y0 2 in
terms of the generalised coordinates q1 = XO, q2 = Yo, q3 = <p we obtain
133
II = "22:: 2:: Cskqsqk , (5.8.8)
s=lk=l
where

Cll = C~l + C~l ' C 22 = C~2 + C~2' }


C33 = b2 C ll + a 2C22 + 2ab (C~2 - C~2)'
(5.8.9)
C 12 = C~2 + C~2' C23 = -bC12 + a (C~2 - C~2)'
C31 = -bCn + a (C~2 - C~2)'

5.8.2 A statically indeterminate system


An example of a statically indeterminate system of rods is shown in Fig.
5.9. All hinges are ball-joints. The natural lengths of the rods are l~, ... , l~
and the direction OkM~ is given by the cosines ak, f3 k , ''tk of the angles
between this direction and the axes Moxyz fixed in space. The potential
energy of the elastic forces under small displacement u = MoM of point
Mo to point M with the coordinates x, y , z is sought.
Under the above displacement the rod rotates through a small angle €g
and its length becomes Lk = OkM = OkM'. Its elongation is fk = Lk -lg
5.8 Calculation of the potential energy for rod structures 235

and neglecting second order terms we can adopt that the vertex angle M'
of the triangle MoM'M is the right angle. Then, denoting the unit vector
of direction OkM~ as e2 we have

and the potential energy becomes

where the axial rigidity of the k - th rod is given by

EkFk
Ck = -Zo-'
k

Therefore

II =

x
1
"2llxyzll C y (5.8.10)
z

where
n n n
Cn = L cka~ C12 = L ck a kf3 k C13 = L Ckak'lk
k=l k=l k=l
n n
c= C22 = L ckf3% C23 = L ckf3 k 'lk . (5.8.11)
k=l k=l
n
C33 = L Ck'l~
k=l

Let us consider now the same system of rods, but under the assumption
that the rod lengths Zk due to assembly differ from the initial lengths Z2.
Denoting as above the elongation of the k - th rod by fk we have

where bk stands for the change in length caused by system assembly, with
positive bk corresponding to the tension. The potential energy of the k - th
rod is equal to

In order to calculate the second component in the above formula we need


an expression for fk up to the second order terms in x, y, z. Since the
236 5. Work and potential energy

coordinates of points Ok and Mare (-Qklk' -f3klk' -'Yk1k) and (x, y, z),
respectively, we obtain

fk Lk -lk = j(x + Qklk)2 + (y + f3k1k)2 + (z + 'Yk 1k)2 -lk


[(1 + 2 QkX + ~:y +'Yk z + x 2 +~; + z2r/2 -1].

Expanding the radical into a series we omit the terms of order higher
than two. The result is

!k Qk X + f3 ky + 'Yk z + 2~k [x 2 (13% + 'Y%) + y2 ('Y~ + Qn +


z2 (Q~ + f3~) - 2Qkf3kxy - 2f3k'YkYz - 2'YkQkZX] .

Cancelling out the term corresponding to the potential energy of the system
in the initial position, we obtain now

x
Y (5.8.12)
z

where

n n n
R1 = L CkOkQk, R2 = L ck Okf3k, R3 = L CkOk'Yk, (5.8.13)
k=1 k=1 k=1

(711 (712 (713


(7 = (721 (722 (723 (5.8.14)
(731 (732 (733
5.8 Calculation of the potential energy for rod structures 237

M
~'VVV'vQ;vvvvirf.---i~~
I_ 2l -I
FIGURE 5.10.

with the entries given by

Cll t Ck [a~ + ~k (,8~ + 10] ,


t
k=l k

C 12 ckak,8k (1 _~k) ,

t
k=l k

C 13
k=l
Ckaklk (1 _~k) k
,

C22 t Ck [,8~ + ~k (r~ + aD] ,


t
k=l k

C 23 Ck,8klk (1- ~k) ,


t
k=l k

C33 Ck [,~ + ~k (a~ + ,8%)] ,


k=l k
C 21 C 12 , C32 = C 23 , C 13 = C31 ·
Notice that the quantity

(5.8.15)

is equal to the initial stress in the k - th rod. Thus R 1, R 2 , R3 represent


the projections of the resultant vector of forces at the joint Mo. Provided
that there is no external force applied to point Mo the initial configuration
is in equilibrium, i.e.

(5.8.16)

The potential energy in then a homogeneous quadratic form, eq. (12).


This form is not necessarily positive definite. The explanation for this is
that the potential energy is evaluated from the equilibrium configuration
and not from the natural one. The potential energy in the equilibrium
configuration is zero and can become negative when the system is displaced
from the equilibrium.
In order to show that this is feasible, let us consider a mass between
two identical initially compressed springs, Fig. 5.10. Denoting the initial
238 5. Work and potential energy

FIGURE 5.11 .

compression of each spring by 8 we have

C¥l 1, = -1, /3 k = 'Yk = 0,


C¥2 81 = 82 = -8,
h b = I, C1 = C2 = C
and the potential energy takes the form

(5.8.17)

5.9 The potential energy of a rod under bending,


torsion and compression
Let us consider the natural configuration of an elastic rod whose axis is a
locus of the centre of inertia of the cross-sections and is a spatial curve,
see Fig. 5.11. The end 0 of the rod is clamped which means that the
end cross-section can not move and rotate. The other end M is mounted
to a rigid body S which the axes M(ry( are fixed to. We look for the
potential energy of the elastic reaction of the rod under the rotation () and
displacement u of the body S. The final position of the trihedron M(ry(
is designated by M' (ry' ('. The study will be limited to the case of small
displacement and rotation, i.e. the projections u, v, w of the displacement
vector u and the projections C¥, /3, 'Y of the rotation vector () on axes M(ry(
will be taken as small quantities so that their products will be neglected.
The rotation about axes M(, Mry, M( through the angles C¥, /3, 'Y makes the
trihedron M(ry( parallel to the trihedron M'(ry'('. The order ofrotation is
immaterial which follows from the commutation of infinitesimal rotations.
Therefore the following six quantities

(5.9.1)
5.9 The potential energy of a rod under bending, torsion and compression 239

-l.

FIGURE 5.12.

are adopted as the generalised coordinates. The corresponding generalised


reactions are the projections of the resultant vector VO and the resultant
moment LO of the elastic reaction in the rod

vt = Ql, v~ = Q2 , vt = Q 3, L~ = Q4, L~ = Q5, L~ = Q6·


(5.9.2)

Here - VO and -Lo represent respectively the resultant vector and the
resultant moment of the forces applied to the rod at cross-section M. A
slow and monotonic increase in these forces makes the natural configuration
coincide with the equilibrium one provided that the rod mass is neglected.
Figure 5.12 shows a free-body diagram of the rod. Let us consider the
part A+O and denote the resultant force and the resultant moment of the
forces in cross-section A as - V and - L. The part M A+ of the rod is in
equilibrium under force - V O and moment -Lo in cross-section M and
V and L in cross-section A_. The equilibrium equations for this part, ex-
pressing the condition of the resultant vectors of all forces and all moments
about point M, are given by

or

(5.9.3)

Here r (s) and s denote the position vector of point A of the rod axis
(the origin at point M) and the curvilinear coordinate along the rod axis,
respectively. The projections of r (s) on axes M ~ 7]( are designated by
~ (s) , 7] (s) , ( (s). We consider the equilibrium configuration of the rod,
and vector r (s) should be determined in this unknown configuration, which
presents a challenging problem in the case of finite displacements. In the
case of small displacements and rotations r (s) is identified with the posi-
tion vector MA of point A in the natural configuration of the rod, then
~ (s) ,7] (s) ,( (s) are prescribed functions of s, the latter being counted
240 5. Work and potential energy

along the axis of the undeformed rod. We will use the axes Axyz with
the origin at point A (the centre of inertia of the cross-section), with the
axes x and y being directed along the principal axes of inertia of the cross-
section, and axis Az along the tangent to the rod axis. The direction cosines
aik of the angles between the axes Axyz and axes M~1]( are prescribed,
too.
Projecting the second equation in (3) on axes M~1]( we obtain

Lr; = L~ -1] (8) vt + ((8) v~, }


L7) = Lg - ( (8) Vr;O + ~ (8) Vc;O, (5.9.4)
Lc; = Lg - ~ (8) V~ + 1] (8) vt
Projections of vector L on axes Axyz which are the bending moments
Lx, Ly and the torque Lz are found with the help of the following coordinate
transformation

Lx = Lr;a11 + L7)a12 + Lc; a l3, }


Ly = Lr;a2I + L7)a22 + Lc;a23, (5.9.5)
Lz = Lr; a 3I + L7) a 32 + Lc;a33.
The potential energy of the rod is composed of the potential energies of
bending and torsion

II = -
1 JI (L2 + -11..
L2 + ~
--'E.
L2) d8 (5.9.6)
2 ABC .
°
Here A = Elx and B = Ely are the bending rigidities which are equal
to products of the Young's modulus E and the moments of inertia of the
cross-section Ix and Iy about the corresponding axes, C is the torsional
rigidity and depends on the geometrical characteristics of the cross-section
and the shear modulus, and l is the rod length. Notice that we neglected the
potential energy of tension which implies that the rod axis is inextensible.
Substituting Lx, Ly,L z from eqs. (5) and (4) into (6) and replacing
L~, ... , Vc;o by the generalised forces (2) we come to the following equation
for II as a quadratic form of the generalised forces

1 66
II = '2 L LtJSkQsQk' (5.9.7)
s=lk=1
where the 21 elements tJ sk of the influence matrix tJ are the definite integrals
of some functions of 8 depending upon the geometric form of the rod axis
in the natural configuration and the rigidities A, B, C. The stiffness matrix
is the inverse of the influence matrix, i.e. c = tJ- I .
To study the structure of the influence matrix more closely we use ma-
trix notation. Let L, Lo, Va,
~ denote the column matrices of projections of
5.9 The potential energy of a rod under bending, torsion and compression 241

the vectors L,Lo, VO,r(s) on axes M~77(, respectively. We denote a skew-


symmetric 3 x 3 matrix accompanying the column-matrix ~ due to the rule
(A.2.3) as ~ and the matrix which makes the axes M~77( parallel to the
axes Axyz as a. Equations (4) and (5) can be cast as follows

(5.9.8)

where L stands for the column-matrix of the projections of vector L on


axes Axyz. Entering the diagonal matrix of the rod compliance
A-I o
(J= o o (5.9.9)
o C- 1

leads to the quadratic form in terms of the projections of vector L on axes


Axyz

(5.9.10)

which is equal to the potential energy of the unit length of the rod axis.
Taking into account that

L' = L~a' + V~~a',


we obtain

Introducing now the following matrices

J~a' (Ja~ds, J, - J
I I I

{31 = - ry = - a (Ja~ds, {32 = a' (Jads (5.9.12)

° ° °
we can rewrite eq. (6) for the potential energy in the form

J
I

II = ~ L' (JLds = ~ (V~{31 Va + V~ry' Lo+ L~ryVo + L~{32Lo), (5.9.13)

°
where the second and the third terms are equal to each other.
In order to derive representation (7) we introduce the 6 x 1 column matrix
of the generalised forces

Q= II f~ II (5.9.14)
242 5. Work and potential energy

and the following 6 x 6 symmetric matrix

(5.9.15)

Because

V~;31 Va + V~I" La + L~I'VO + L~;32Lo


we obtain by virtue of (13)

(5.9.16)

which is the desired relationship (7). Formulae (12) and (15) show how to
calculate the influence matrix ;3.
We proceed now to some special cases.

5.9.1 Plain curve


The calculation is simplified when the curve axis in the natural configura-
tion is a plain curve, one principal axis of inertia of the cross-sections lying
in the plane of the rod and the other being perpendicular to this plane.
Changing slightly the above notation we make the rod plane coincident
with plane M~TJ(. Axis x of the basis system Axyz is directed along the
tangent to the rod axis whereas axes y and z are directed along the principal
axes of inertia of the cross-section, namely y in the plane M ~TJ and z parallel
to axis (. The equations of the curve are given by

and the table of the direction cosines rY-ik is as follows

I I ~ I TJ I ( II
x ~ r/ 0
,
y -TJ' ~ 0
z 0 0 1

where a prime denotes differentiation with respect to s.


Equations (4) and (5) are split into two groups. The first one operates
with the quantities describing the bending in the rod plane, i.e. V,?, V~ and
La(

(5.9.17)
5.9 The potential energy of a rod under bending, torsion and compression 243

Changing notation (2) we introduce the generalised forces of elastic re-


actions and the corresponding generalised coordinates

L,
o--Q'
3, q1 = U, q2 = v, q3 = "/. (5.9.18)

The second group describes rod bending in which the points of the rod
axis leave plane M~'T/, bending being accompanied by torsion. This group
operates with V,o, L~, Lg. By virtue of eqs. (4) and (5) we have

Lx = Lee + L1)'T/',
(5.9.19)
Ly = -Le'T/' + L1)e,

where

LOe-
- Q6', a = q6. (5.9.20)

The expression for the potential energy is also split into two terms

Denoting

(5.9.21)

we obtain the following expression for the elements of the influence matrix
(3(1)

(3(1) -
11 -
J
I

'T/ 2ds
El z '
(3 (1) -
12 -
- J
I
~'T/ds
Elz '
(3 (1) -
13 -
J
I
'T/ds
El z '

J° J° J°
I I
(5.9.22)
I 2
(3 (1) - ~ ds (3 (1) - - ~ds (3 (1) - ds
22 - El z ' 23 - El z ' 33 - El z '
° ° °
244 5. Work and potential energy

Expressions for (3(2) are more complicated

(3 (2) = 11 [(~1]1 _1]e)2 + (~e + 1]1]/)2] d


11 CBS,

1[(~1]' - 1]e) 1]' + (~e + 1]1]IH/]


o
1
(3 (2) = d
12 CBS,

1[(~1]1
o
1
(3 (2) = (~e + 1]1]') 1]1]
d
-1]e) (
-
(5.9.23)
13 CBS,
o
(2)
(322 =
11 ((2Ii + 1]/2) Cds, (323 =
(1) 11 (1C - 1) 1 1 B ~ 1] ds,
o 0

(1) _
(333 - 11 (CC/2 + Ii1]/2 )
o
<"
ds.

As 1] = 0 and e = 1 for the rod which is straight in the natural configuration


we obtain

(3 (1) -
22 -
1 1
~ ds
2

Elz '
(3 (1) -
23 -
-1 1
~ds
El z '
(3 (1) -
33 -
1 1
ds
Elz
o o
1 1 1
o
1 2 1 1

(3 (2) - ~ d~ (3 (2) - ~d~ (3 (2) - d~ (5.9.24)


11 - El' 12 - El' 22 - El'
y
o o y

1c'
y
o
1

(3 (2) - d~
33 -
o

The other elements are identically equal to zero. If we assume that the rod
is extensible it can be shown that

(3 (1) -
11 -
1 1

EF'
d~ (5.9.25)
o
5.9 The potential energy of a rod under bending, torsion and compression 245

~
IJ
n t'
oX z

FIGURE 5.13.

where F denotes the cross-sectional area of the rod. The potential energy
can be expressed as the sum of four components

1 (1) 2
III = 2'(311 QI'

II2 = 2'1 ( (322 Q2 + 2(323 Q2Q3 + (333 Q 3


(1) 2 (1) (1) 2)
,
(5.9.26)
1 ( (2) 2
II3 =2 (311 Q4 + 2(312 Q4Q5 + (322 Q 5
(2) (2) 2)
,

II - ~(3(2)Q2
4 - 2 33 6'

Here II2 and II3 correspond to bending in the planes ~ry and (ry, respectively,
IIIand II4 correspond to the tension and torsion, respectively.

5.9. 2 Helical spring


Here the natural configuration of the rod axis is a helix on the surface of a
circular cylinder of radius a. Let D( be the cylinder axis and axes D~ , Dry
be perpendicular to this axis, see Fig. 5.13. The equations of the axial line
of the spring are as follows
c scosa . scosa
<" = acos---, ry = aSln---, ( = ssina, (5.9.27)
a a
where a is the helix angle and s is the curvilinear coordinate along the rod
axis. The lower end of the spring is clamped to a rigid plate S lying in the
plane D~ry and the upper end is clamped, too. The spring is made of a wire
of length l with a circular cross-section of radius r.
The plate is assumed to be subject to a force V O applied at point 15 and a
moment LO. The displacement of point 0 and the rotation angle of the plate
are sought under the assumption of small displacements and rotations. The
axis Az coincides with the unit vector T of the tangent to the helix. Axes
246 5. Work and potential energy

Ax and Ay can be directed along the normal n and the binormal b since
any two mutually orthogonal directions of the circular cross-section can be
taken as principal axes of inertia. Denoting the position vector of point A
as r we have, due to eq. (2.18.2),
dr d2 r
T = ds' n = p ds 2 ' b =T X n,
where p is the radius of curvature. We have

Tt; = -cosasm---,
. scosa
a
TTl =
scosa
cosacos---,
a
T, = sin a
and furthermore that
nt; = - -
p
a
cos
2
a
a
scosa
cos - - - ,
TI
P 2 • scosa
n = --cos aSln---
a a
n, =0,

where p = a/ cos2 a because of the unit length of n. The table of direction


cosines is

II II ( II
x
y
-coscp
sin a sin cp
- sincp
- sinacoscp °cosa ,
z - cosasincp cos a cos cp sin a

where cp = scosa/a. Projections of the moment in cross-section A on axes


Axyz are determined by means of eqs. (4), (5) and (27). Although the
calculation presents no problem, we restrict our consideration to the case
of the axial force and the axial moment
Vzo = Ql, Lg = Q2.
The corresponding generalised coordinates
ql = W, q2 = I

are the displacement of plate S along axis O( and the rotation angle about
this axis, respectively.
The calculation yields
Lt; -aV,osincp, LTI = aV,ocoscp, L, = Lg,
Lx 0, Ly = -aV,o sin a + Lg cos a, Lz = aV,o cos a + Lg sina
and
(3 _[2
(
sma
. 2 cosa
2 )
a B + C '

(1 1).
11 -

(312=[a C- B smacosa, (5.9.28)

(3
22-
- (cos 2
B
a+ a) sin 2
C .
5.10 Power 247

The inverse matrix is


1 .
Cll = la12 (C cos 2 a + B sm
. 2 )
a C12 = la (B-C)smacosa
c=
1 .
C22 = l (Csm 2 a + Bcos 2 a)
(5.9.29)
Therefore the generalised elastic reactions of the spring are equal to

Q1 Ccos2a+Bsin2a B-C. }
= - la 2 W - -l-a-ry sm a cos a,
(5.9.30)
B - C C sin 2 a + B cos 2 a
Q2 = -~wcosasina - l T

For a wire with circular cross-section


7fEr 4 7fEr 4
B=--
4 ' C= 4(1+v)'
where v is Poisson's ratio.
Applying only an axial force causes not only axial displacement w but
also rotation of the plate through the angle
C - B w . vsinacosa w
ry = 2 sm a cos a = - -. (5.9.31 )
C sin a + B cos 2 a a 1 + v cos 2 a a
Similarly, applying only a torque results not only in rotation ry but also
in axial displacement
(C - B) sin a cos a v sin a cos a
w = ary = - ary. (5.9.32)
2
C cos 2 a + B sin a 1 + v sin 2 a
These effects are proportional to sin a, that is they are small for small helix
angle a.

5.10 Power
The power of actual motion of the system is the sum of scalar products of
the force vectors and the velocities of the points where the forces are applied
to, i.e.

(5.10.1)

or by definition (1.3) of the generalised forces


n N {}
N=2::QsQS+2:: F i. ~i. (5.10.2)
s=l i=l
248 5. Work and potential energy

The first term referred to as the virtual power


n
(5.10.3)

is obtained by replacing the actual velocity Vi in the definition for the


power by the virtual velocity V;.
The concepts of the power of actual motion of the system and the virtual
power coincide when the system is subject to stationary constraints.
An expression for the power in terms of quasi-velocities is obtained by
means of eqs. (1.5.8) and (1.8)
n n n

N = LQs Lbsrwr = LPrWr. (5.10.4)


s=l r=l r=l
While deriving this result we assumed that the quasi-velocities are related
to the generalised velocities by linear relationships (1.5.1).
For instance, by virtue of eq. (5.2), the power of the forces acting on the
rigid body points can be represented in either of two ways: in terms of the
generalised forces and generalised velocities such that

(5.10.5)
or in terms of the quasi-velocities and generalised forces such that
(5.10.6)

5.11 The dissipation function


Dissipative force resists the motion since it acts in the direction opposite
to that of the velocity. We consider the resisting forces which can be rep-
resented in the form

Fi = -kddvi) -.!:.. (i = 1, ... ,N), (5.11.1)
Vi
where Vi is the absolute value of point M i , and k i and fi (Vi) denote positive
functions of the generalised coordinates and velocities Vi, respectively.
Due to eqs. (1.3) and (1.3.5) the generalised forces corresponding to the
forces (1) are defined by the equalities
N N
Qs = -
'"'
L kdi (Vi) - . -ari
Vi
= -
'"'
L kdi (Vi) -Vi . -
aVi
..
i=l Vi aqs i=l Vi aqs
Noticing that
5.11 The dissipation function 249

we find that

(s=I, ... ,n),

(5.11.2)

where <I> denotes the dissipation function

k JIi (u) duo


N

?=
Vi

<I> = i (5.11.3)
t=1 0

Clearly, <I> ;::: 0 since all the integrands are positive.


The generalised forces corresponding to the quasi-coordinates are, due
to eq. (1.8), equal to

n n 8<1> 8<1>
Ps = LbrsQr = - Lbrs~ = - >lw (s = 1, ... ,n), (5.11.4)
r=1 r=1 uqr u s

where Ws denote quasi-velocities.


The dissipation function was introduced by Rayleigh in his classical trea-
tise [75] for resisting forces which are linear in velocities. In the present book
this idea is generalised to more general resisting forces.
When the dependence of the resisting forces on velocity is given by

(5.11.5)

the dissipation function

<I> = _1_ 'N" k. ~+1 = _1_ 'N" k. ' n" 8ri . I Im


+1 (5.11.6)
m
+1 ~
i=1
tVt
m
+
1 ~ t ~ 8 qs
i=1 s=1 qs

is a homogeneous function of degree (m + 1) in the generalised coordinates.


Using Euler's theorem on homogeneous functions (4.1.12) we can easily
relate the dissipation function to the power of the dissipative forces

n n 8<1>
N = L Qsqs = - L ~qs = - (m + 1) <1>. (5.11.7)
s=1 s=1 qs
In passing we note that m = 0 corresponds to Coulomb's friction, m = 1
to the dissipative forces of the Rayleigh type, i.e. linear in velocities, and
m = 2 to square-law resisting forces.
While deriving equations for the generalised forces due to (2) it is nec-
essary to bear in mind that one should differentiate expressions containing
250 5. Work and potential energy

absolute value of the generalised velocities. For instance in the case when
the resisting force is proportional to an even power of the velocity and
n=1

<p = _k_I·12S+l
2s +1 q ,

where signq = +1 if q > 0 and signq = -1 if q < O.


Let us consider the motion of the system relative to moving axes Ox' y' z'.
Here we assume that expressions for the position vectors r~ of the system
points with respect to the pole a do not contain time explicitly. The re-
sisting forces acting on the system points are determined by their velocities
relative to the environment which do not coincide with the velocities v~
relative to the basis Ox' y' z' .
For example, let us consider motion of a mathematical pendulum whose
velocity of the attachment point is va relative to the earth. Let the motion
take place in the air flow having velocity V and the pendulum velocity
relative to the axes moving together with the attachment point be v'. The
absolute velocity of the pendulum (the velocity relative to the earth) is
va + v' whereas its velocity relative to the flow is va + v' - V. The latter
expression for the velocity must appear in the equation for the resisting
force
, va +v'- V
F = -kf(lvo +v - VI) I
vo+v ' - VI

When, for instance, V = 0, that is the air does not move, the resisting
force is defined by the absolute velocity of the point which coincides with
the velocity relative to the air. When va = v' = 0 the point is subject to
the force

F = kf(V) ~
directed along the air flow.
Returning to the general definition, we rewrite equation (1) for the re-
sisting force in the form

(5.11.8)

where vi denotes the geometric difference of the vectors of translation ve-


locity and absolute velocity of the environment medium Vi at that point of
the moving system where point Mi takes place instantaneously. Vectors vi
5.11 The dissipation function 251

can depend on time t and the generalised coordinates determining the posi-
tion of points referring to axes Ox' y' z' but not on the generalised velocities
qs. Vectors v~ are linear forms in the generalised velocities.
In particular, vi can be equal to zero and then the resisting forces will
depend only on the relative velocities. It can occur in those cases when
the medium in which the motion takes place moves together with the axes
Ox'y'z'.
The dissipation function is introduced by formula (3). Equation (2),
which yields the generalised forces, remains valid also. It follows from the
fact that the expression for the generalised force can be taken in the form
(1.11). In addition to this

ar~ 8v~
aqs = aqs = aqs
a (* ')
vi +vi ,

because vi is independent of the generalised velocities. Repeating the


derivation of formula (3) we obtain

(5.11.9)

where now
Iv:+v:1
Jf
o
(u)du. (5.11.10)

Provided that dependence (5) holds, the dissipation function remains a


homogeneous function of qs only if vi = o. This is, for example, the case
when axes Ox'y' z' move together with the medium.
The power of the dissipative forces in actual motion is
N N N

N = l:Fi· Vi = l:Fi· v~ + l:Fi· vie,


i=l i=l i=l
where Vi and Vie denote vectors of absolute and translation velocities,
respectively. The first term describes the virtual power given by eq. (5.10).
We obtain

(5.11.11)

where Qs is given by formula (9).


252 5. Work and potential energy

Oo:::::--- --y'
0/1

FIGURE 5.14.

5.12 Examples of the calculation of the dissipation


function
5.12.1 Double mathematical pendulum with a square-law
resisting force
The aim here is to obtain expressions for the dissipation function and the
generalised resisting forces for the double mathematical pendulum shown in
Fig. 5.14. The suspension point 0 moves with a velocity Vo in the moveless
air, and the resisting force of the air is taken to be proportional to the
square of the velocity relative to the air.
The system motion will be described relative to axes Ox' y' moving trans-
lationally with velocity Vo. The squares of the absolute velocities of the
points are given by
v~ Iv~ + vol 2 = (vocoso: - h<Pl sin <Pl)2 + (vosino: + h<Pl cos <Pl)2
= v5 + l~<pi + 2voh <PI sin (0: - <PI) , (5.12.1)

v22 I'
v2+ v O12 = (vocoso:- l" · ·)2 +
l<Pl sm <Pl- l 2<P2sm<P2
(vo sin 0: + h <PI cos <PI + b<P2 cos <P2)2 (5.12.2)
v~ + l~<p~ + 2vol2<P2 sin (0: - <P2) + 2hl2<Pl<P2COS (<PI - <P2)'
Since
u

Jf(U)du=~u3,
o
the dissipation function, due to eq. (11.10), is as follows

<P 1 (3
=3 k 1 v 1 +k2 V 23) . (5.12.3)
5.12 Examples of the calculation of the dissipation function 253

The generalised forces are equal to

-klVI [l~01 +voh sin (a-'Pl)] - }


k2V2 [l~01 + voh sin (a - 'PI) + hl202 cos ('PI - 'P2)]
-k2V2 [l~02 + VOl2 sin (a - 'P2) + hl201 cos ('PI - 'P2)] .
(5.12.4)

In the relative equilibrium 01


form
°
= 02 = and the generalised forces take the

Q~ = (kl + k2) v5h sin ('P~ - a), Qg = k2V5l2 sin ('Pg - a) ,


which is easy to prove by direct calculation of the elementary work of forces
-kiVOVO (i = 1,2) due to virtual displacements 8rl and 8r2. The elementary
work of the gravity force in these virtual displacements equals

- PI h sin 'P~ 8'Pl - P2 (h sin 'P~ 8'Pl + l2 sin 'Pg8'P2) ,


and the generalised forces due to gravity are

Qi = -(PI +P2)hsin'P~, Q2 = -P2l2sin'Pg.


From the equilibrium equations (see Sec. 6.5)

Ql + Qi = 0, Q2 + Q2= 0,

we obtain the values of 'P~ and 'P~


o PI +P2 o P2
cot 'PI = cot a - (k k) 2· , cot 'P2 = cot a - k 2· .
1 + 2 Vo sma 2Vo sma
(5.12.5)

Expressions (4) for the generalised forces turn out to be very complicated.
They can be considerably simplified if small oscillations of the system about
the position of the relative equilibrium are considered. In this case we adopt
the angles €1 = 'PI - 'P~ and €2 = 'P2 - 'P~ as well as the angular velocities
El and E2 to be small quantities. Then, neglecting the products and the
squares of these quantities and taking for simplicity a = 7f /2, we obtain

VI = Vo + hEI cos 'P~, V2 = Vo + hEI cos 'P~ + bE2 cos 'Pg


and furthermore
o
tan'Pl = -
kl
p
+ p.
k2 2
Vo,
1 + 2

In view of these equalities we have

kl +
cos 'PI = cos 'PIo ( 1 + €1 PI k2 Vo2) ,
+ P 2
254 5. Work and potential energy

@ Wz

o (S)
u"
-x

FIGURE 5.15.

0
cos ( 'PI -
0) 0 0 [
= cos 'PI cos 'P2 1 + (PI
(kI + Pk 2)) k2
+ 4]
'P2 2 P2 Vo .

Inserting these equations into eq. (4) yields

(5.12.6)

These formulae are meaningful only for non-zero and sufficiently large Vo
since if Vo = 0 the expressions for vr
and v~ become quadratic forms of EI
and E2.

5.12.2 Coulomb's friction


Let us consider a rigid plate S pressed against a motionless rough plane by
force G, Fig. 5.15. The power needed for rotation of the plate with angular
velocity W z about the pole 0 which moves with velocity Vo is sought here.
The force of Coulomb's friction acting on the element do of the contact
surface is equal to
v
-fpdo-,
v
where f is the coefficient of friction which is assumed to be independent of
velocity, and p is the pressure which is assumed to be uniformly distributed.
5.12 Examples of the calculation of the dissipation function 255

The summation in eq. (11.6) is replaced by integration over the contact


surface S, i.e.

<I? = fp 11s
vdo. (5.12.7)

We have
v = Vo +w x r

or in projections on axes Oxy bound to the plate

Therefore

fp 11 V
s
(vOx - Wzy)2 + (VOy + Wzx)2do

fp IWzl 11 s
V(xp - x)2 + (yp - y)2 do . (5.12.8)

Here
VOy VOx
xp = --, yp=- (5.12.9)
Wz Wz

are the coordinates of the instantaneous centre of velocities. Denoting the


distance between the above centre and the element do by r we can write
(5.12.10)
where r* has the dimension of length and is equal to

r* = ~ 11s
rdo. (5.12.11)

Taking the instantaneous centre of velocity as the origin of the polar


coordinate system (r, >.) with the polar axis OP we have do = r dr d>..
Hence, if the instantaneous centre is inside of S then as Fig. 5.16a shows

1
211'

r* = 3~ r3 (>.) d>', (5.12.12)


o
where r (>.) = PM, M being a generic point on the border. If it is outside
S, then, as Fig. 5.16b displays,

1[r~
).,2

r* = 3~ (>.) - d (>')] d>', (5.12.13)


).,1
256 5. Work and potential energy

FIGURE 5.16.

where r2 (A) = PN, rl (A) = PM and the angles Aland A2 are shown in
Fig.5.16b.
Thus the value of r* depends on the form of the plate and the position
of the instantaneous centre of velocity. For a circular plate of radius a we
have in the first case, cf. Fig. 5.17a,

r3 = (RcosA+Ja2-R2sin2A)3 =R3cos3A+(a2_R2sin2A)3/2

+3R2 cos 2 AJa2 - R2 sin 2 A + 3Rcos A (a 2 - R2 sin 2 A) ,


where R = OP. As the integrals corresponding to the first and the last
terms equal zero, we obtain

(5.12.14)

where K and E are the complete elliptic integrals of the first and second
kind with the modulus

O<k=R<l.
- a-
In the second case, see Fig. 5.17b, we have

r~-rr = (r2 - rl) [4 (rl + r2; rl) 2 - r l r 2] = MN [4(PQ)2 _ (PT)2]

= 2Ja2 - R2 sin2 A (4R2 cos 2 A - R2 + a2)


and furthermore

Substituting
R sin A = a sin e
5.12 Examples of the calculation of the dissipation function 257

FIGURE 5.17.

yields

We have

r* a (a) - (1 -a a2 ) K (a)]
3 2 2
- =4R
-- [ ( 1+7- ) E - - ) (1+3- -
a 91f a3 R2 R R2 R2 R'
(5.12.15)

where K and E are the complete elliptic integrals of the first and second
kind with the modulus

O<k=R<1.
- a-
In eqs. (14) and (15)

(5.12.16)

The limiting case Vo = 0, which describes rotation about the fixed axis
passing through the circle centre 0, is obtained from eq. (14) by setting
R = O. In this case

E (0) = K (0) =
1
21f, r* =-a
2
3 '
(5.12.17)

which can be proved easily.


258 5. Work and potential energy

The second limiting case is the translatory motion. Recalling the expan-
sions

we obtain from eq. (15) by taking the limit that

(r*)
R R--->oo
=~{~[(1+7k2)(I_k2 + ... )_
9 k 4

(1+2k 2 + ... ) (1+ 4 2


k )]}
k2-->O
= 1
and hence

(5.12.18)

which is the aim of the analysis.


In the case when a circular disc is rolling without slip we take R = a. It
follows from eqs. (14) and (15) that

32
r* =-
32a
il> = - fGa IWzl = 1.13fGvo,
97f ' 97f

where Vo denotes velocity of the disc centre. This expression is easy to


derive directly from eq. (11) by setting r(A) = 2acosA and taking the
lower and upper limits of the integral as -7f /2 and 7f /2, respectively.
The generalised forces of friction corresponding to the quasi-velocities VOx
and VOy which are the projections of the resultant of the frictional forces on
axes Ox and Oy fixed in the plate are, due to eqs. (11.4) and (16), equal
to

ail> oR ail> VOx ail> ok2


Px ----
oVox oRovo x -IWzl Vo ok 2 oR'
VOy ail> ok2
Py
-IWzl Vo Ok2 oR·
Applying the relationships

oK 1 E - (1- k 2 ) K oE 1K - E
Ok2 ="2 (1- k 2) k 2 Ok2 -"2~

we obtain

Px__
-
vOxp
, (5.12.19)
Vo
5.12 Examples of the calculation of the dissipation function 259

Here

R:S; a,

a :s; R.
(5.12.20)

A simple way to find the generalised force Pw , which is a torque, is to


use the formula for the power

that is

Pw = - - Vo = ( PR- - <I> ) slgnw


<I> +P- • z.
Wz Wz IWzl
Then we find
.
Pw slgnw 4fG
z = --a x (5.12.21)
97l'

The limiting cases are:


a) rotation about a fixed axis passing through the centre of the plate

p=o, (5.12.22)

b) translatory motion (R --+ 00)

P= fG, Pw =0, (5.12.23)

c) rotation without slip (R = a)

Pw = - 8fG asignw z , (5.12.24)


97l'
which is equivalent to a force of magnitude P applied at a distance of
4a/3 from the instantaneous centre of velocities, the force direction being
opposite to the velocity of the point at which the force is applied.
Calculation of <I> for a polygonal plate is cumbersome but elementary.
Figure 5.18 shows a rectangular plate having sides 2a and 2b. Considering
260 5. Work and potential energy

1-+--+- 20

FIGURE 5.18.

the simplest case, that is rotation of a fixed axis passing through the centre
of symmetry 0, we obtain by eq. (12)

where
b
tanA2 = -.
a
Here similar to any polygonal plate the problem is reduced to integration
of expressions containing the following integrals

J dA
- - = -1 [tanA
cos3 A
--
2 cos A
+ In 1 (7r + -2A) I] .
tan -
4
In the case under consideration we obtain

<P = ~fGlwzl{Ja2+b2+ (5.12.25)

2~b [b 3
ln Itan ( ~ + ~l ) 1 + a 3 ln Itan (~ + ~2 ) I] } .

5.13 Aerodynamic resisting force


Study of the forces acting on a body (shell, airplane) is the subject of
theoretical and experimental aerodynamics. Consideration of this topic in
a book on analytical mechanics is possible only in general terms and is
aimed at giving an insight into the character and difficulty of the mechanical
problems arising from taking account of these forces.
Following [68J we consider a system of aerodynamic forces acting on a
spinning shell moving through the air at rest. The shell motion with respect
5.13 Aerodynamic resisting force 261

to the earth is described by the velocity vector v of the pole 0 and the
angular velocity vector w. It is common practice in ballistics to take the
centre of inertia of the shell as the pole. However this choice is illogical
as the position of the centre of inertia is defined by the mass distribution
within the shell whereas the aerodynamic forces are conditioned by the
geometrical form of the surface of revolution bounding the shell body. For
this reason the pole which is the origin of the axes Oxyz fixed in the shell
is taken at the centre of the shell volume lying on the shell axis Oz. In
principle, any point on the shell axis could be taken as the pole because
the aim of the forthcoming analysis is to derive such dependences of the
resultant vector F and the resultant moment rno on vectors v and w which
is indifferent to the choice of the pole.
Projections of v and w on the shell axis are denoted by V3 and W3,
respectively. It is presumed that they are much greater than the magnitudes
v* and w* of the components v* and w* lying in plane Oxy.
The vectors F and rno are also represented in the form

rn O = rn o* +km 3o , (15.13.1)

where the transversal force F * and the transversal moment rn~ are respec-
tively the components of F and rno in the plane Oxy and k denotes the
unit vector of axis Oz.
The axial force F3 and the axial moment mf? are assumed to have the
form

(15.13.2)

where p denotes the air density, a the shell radius and hand 93 are non-
dimensional aerodynamical functions determined experimentally. They de-
pend upon the non-dimensional quantities V3/C and aW3/c, c being the local
velocity of sound, and possibly on the Reynolds number. The other aero-
dynamical functions designated by letters f and 9 with the corresponding
subscripts are assumed to be dependent on the above quantities.
The transversal components of the force and the moment are taken to
depend linearly on v* and w*' the position of axes Ox and Oy does not
influence these dependences due to the axial symmetry. With this in view,
the most general representations are as follows

F* = pa 2 (-alv* + b1k x v* + CIW* + d1k x w*), }


(15.13.3)
rn~ = pa3 (-a~v* - b~k x v* - c~w* + d~k x w*),

implying that either of the above vectors consists of four components. Di-
rections of two components are defined by the directions of v * and w*' while
those of the others by the directions of the vectors k x v * and k x w* which
are perpendicular to the above directions and lying in the plane Oxy.
262 5. Work and potential energy

Components blk x v* and ClW* of the transversal force and -aiv* and
di k x w* of the transversal moments characterise the so-called Magnus
effect which appears due to rotation of the shell about axis Oz. The coef-
ficients bl , Cl, ai, di are therefore taken to be proportional to W3 whereas
the other coefficients are taken to be proportional to V3. Thus we have

al = flV3, bl = f2aw3, Cl = gl a2w 3, d l = g2 av3,


ai = ff aw3, bi = f~V3, ci = gi av3, di = g~a2w3.

The expressions for the transversal force and moment take the form

F* = pa 2 (-iIV3V* + haw3k x v* + gla2w3w* + g2av3k x w*), }


m~ = pa 3 (- ff aW3v * - f~V3k x v * - gi aV3W* + g~a2w3k x w*) ,
(5.13.4)

respectively. As pointed out, the ten non-dimensional aerodynamical func-


tions

(5.13.5)

appearing in formulae (2) and (4) depend on V3/C and aW3/c. The signs
are chosen so that the coefficients iI, g2, f~, gi which do not depend On the
Magnus effect are positive when the centre of pressure is located ahead of
the centre of shell volume. The other signs are chosen so that the other
coefficients are positive.
Calculating the aerodynamical functions related to the inertia centre C
of the shell one should bear in mind that the resultant vector F does not
change, while the resultant moment m G is determined by the relationship

(5.13.6)

where ek = CO denotes the position vector of point 0 with the origin at


the inertia centre C. Force F and moment m G should now be expressed in
terms of the velocity of the centre of inertia

(5.13.7)

and vector w which is independent of the pole choice. Similarly to eq. (1)
we have

mG m:+mfk=m~+m~k+ekxF*,
F F: + Ffk = F * + F3k,
from which it follows that eq. (2) and thus, the aerodynamic functions h
and g3 remain unchanged under change of the pole, i.e.
G_
ff=h, g3 - g3· (5.13.8)
5.13 Aerodynamic resisting force 263

Now substituting expression (7) for v * into the equation for F *, we arrive
at the following result

F~ F*=pa2[-iIv3(v~+ekxw*)+
12aw3k x (v~ + ek x w*) + 91a2w3w* + 92av3k x w*] ,

which is reduced to the form

F * = pa 2 (fe
- 1 V3 V e
* + fe2 aW3 k x v*e + 91e a2W3W* + 92e aV3 k x w* ) ,
(5.13.9)

where
e e e e
ff =iI, ff=12, 91 = 91 - -
a
12, 92 = 92 - -iI·
a
(5.13.10)

We also have

m~ pa 3 [- f{aw3 (v~+ ek x w*) - f~V3k x (v~ + ek x w*)-


9~ aV3w* + 92a2w3k x w*] + pa 2ek x

( - f~ V3V~ + 12aw3k x v~ + 9f a2w3w* + 9f aV3k x w* )

and this expression can be simplified, to give

m*e = pa3 (fie


- 1 aW3v e
* - fie
2 V3 k X V
e
* - 91Ie aV3w* + 92Ie a 2 w3 k x w* ) ,
(5.13.11)

where

~ I e ~ I e I e
f1 =f1+-12, 91 =91--(f2-92)-2iI,
a a a2 }
(5.13.12)
2
f 2ie = f'2 + -e f 1, 92Ie = 92I - -e (f'1 - e f 2·
91 ) - 2
a a a

The system of forces and moments given by eqs. (2) and (4) or eqs. (2),
(9) and (11) is the most general provided that the assumption of linear
dependence of the transverse force and moment on the transverse com-
ponents of the vectors of velocity and angular velocity is adopted. The
practical application however is hardly possible and expeditious because
of the complexity and difficulties of the experimental determination of all
ten aerodynamical functions (5). The Magnus effect is primarily neglected,
that is, the aerodynamical forces on the velocity and angular velocity of the
shell are independent of the position of the pole. Indeed, if the coefficients
12,91,f{,9~ are set to zero, then, as follows from eqs. (10) and (12), all
the coefficients with the superscript C vanish, too. Therefore, the above
264 5. Work and potential energy

IJ

FIGURE 5.19.

assumption yields the following expressions for the force and the moment
related to an arbitrary pole

F pa 2 (-khv§ - !Iv3V* + g2av3k x w*) ,


= }
(5.13.13)
rno = -pa3 (kg3aw3v3 + f~V3 k x v* + g~av3w*),

provided that v * denotes the transverse component of velocity of this pole


and the aerodynamical functions are related to this quantity.
The results of test firing are contained in an important investigation on
the aerodynamics of spinning shells [25]. The system of forces and moments
whose existence was experimentally confirmed by the authors is shown in
Fig. 5.19. The figure displays the velocity vector v of the centre of inertia,
the unit vector k of the shell axis comprising angle <5 with the velocity
vector, the unit vector i 1 having the direction of the transversal component
v * of the velocity vector v and the unit vector i2 = k x h. Angle <5 is
assumed to be small, i.e. cos <5 :::::: 1, sin <5 :::::: 0, so that the velocity value can
be identified with V3. The vector w* of the transverse angular velocity of
the shell lies in the plane of vectors i 1, i 2 .
The aerodynamic forces reduce to the forces Rand L and the moments
M and H.
The force R represents the head resistance and is taken to be equal to
R = _pa 2fRV3V = _pa 2 (JRkv§ + fRV3V*) , (5.13.14)

as v = kV3 + V *.
The magnitude of the lateral force is
L = pa 2hv 2<5.
Due to Fig. 5.19 it can be cast in the form
L =pa2h (-v2 <5i 1 + v 2<5 2k) :::::: _pa 2h V3V* , (5.13.15)
5.13 Aerodynamic resisting force 265

where we neglected the term proportional to 82 and took

The tilting moment M and the retarding moment H are given by

(5.13.16)

Expressions for the resultant force F and the resultant moment m C of


the aerodynamic forces about the centre of inertia take the form

F = _pa 2 [fRkv~ + (fR + h) V3 V *] , }


(5.13.17)
m C = -pa3 (fMV3k x v* + fHav3w*).

These relationships enable one to link the aerodynamic functions in eq.


(13) with the corresponding coefficients

(5.13.18)

Additionally we take that

g3 = 0, gf = o. (5.13.19)

Though the system of forces (17) can be sufficiently precise in practice,


it is not consistent since the latter equality in eq. (19) is not invariant.
Adopting absence of term V3k x v in the expression for F when the pole is
chosen at the centre of inertia, we will immediately obtain a similar term
when another pole is taken.
6
The fundamental equation of
dynamics. Analytical statics

6.1 Lagrange's equations of the first kind


Constraint forces are the forces exerted at the points in the system when
the constraints are mentally removed. Introducing into consideration the
constraint forces we distinguish between two categories of forces acting at
the points within the system, namely the constraint forces and the active
(or prescribed) forces. The resultant of the constraint forces exerted at
point Mi is denoted by Ri whereas that of the active forces is denoted by
Fi·
Introducing the constraint forces enables us to write the differential equa-
tion for any particle in the form of Newton's second law

(6.1.1)

which is the differential equation of a constraint-free particle. The above is


the essence of the principle of constraint release. In eq. (1) mi denotes the
mass of particle Mi and withe acceleration vector referring to an inertial
coordinate basis Oxyz

(6.1.2)

Projecting eq. (1) on the coordinate axes and applying the notation of eq.
(1.2.1) we have

mv~v = Fv + Rv (v = 1, ... ,3N), (6.1.3)


268 6. The fundamental equation of dynamics. Analytical statics

where m3s-2 = m3s-1 = m3s for s = 1, ... ,N. In addition to these 3N


differential equations we have r equations for the holonomic constraints

(6.1.4)

and r' equations for the non-holonomic constraints


3N
L lkv~v + gk = 0 (k = 1, ... ,r'). (6.1.5)
v=l

The unknown variables in eqs. (3)-(5) are the 3N coordinates ~l' , ... '~3N
and the same number of constraint forces. The total number 6N of the
unknown variables exceeds the number of the equations 3N + r' + r by
3N - r' - r, which is the number of degrees of freedom, i.e. the problem is
not indeterminate. At this point it is necessary to make some assumptions
about the character of the constraints.
The elementary work of the constraint forces Ri due to virtual displace-
ments of the particles of the system from the positions under consideration
at time t is given by
N 3N
8'W = l:Ri' 8ri = LRv8~v' (6.1.6)
i=l v=l
The constraint equation holds also in the varied state, that is along with
eq. (4) the following equation holds, too

This means that the variation 8 fk of function fk must vanish, i.e.

8fk
~ Ofk ~ (Ofk
6 !'lC 8~v = 6 ~8xi
Ofk
+ ~8Yi Ofk)
+ ~8zi
v=l U<,v i=l uX, uy, uz,
N
Lgraddk·8ri = 0 (k = 1, ... ,r). (6.1. 7)
i=l

Equalities (5) imposes additional r' conditions on variations 8~v

(6.1.8)

Here aki, bki , Cki denote the projections of the vector eki on axes Oxyz.
Let us prove that the sufficient condition for zero elementary work of the
constraint forces is that these forces are cast as linear forms of the gradient
6.1 Lagrange's equations of the first kind 269

of the constraints and vectors eki, i.e.


r r'
Ri = LAkgradifk + LA~eki (i = 1, ... ,N) (6.1.9)
k=l k=l
or in the equivalent form

r Ofk r' ,
Rv = L Ak o~ +L Aklkv (v = 1, ... ,3N). (6.1.10)
k=l v k=l
Coefficients Ak and A~ are referred to as the Lagrange multipliers or the
multipliers of holonomic and non-holonomic constraints. The number of
equations for each type of constraint coincide with the number of the cor-
responding constraint equations.
The sufficient condition can be proved directly. Substituting expressions
for the constraint forces (10) into the elementary work (6) and taking into
account eqs. (7) and (8) we have

which completes the proof.


On the other hand, provided that the sum of the elementary work of the
constraint forces is equal to zero, one can always choose r + r' coefficients
Ak and A~ so that the 3N quantities Rv are represented by linear forms
(10). Indeed, multiplying each of equalities (7) and (8) by -Ak and -A~,
respectively, and summing up the products obtained with eq. (6) we arrive
at the equality

(6.1.11)

We take now the r + r' coefficients Ak and A~ so that r + r' terms of the
sum (11) is equal to zero. Without loss of generality we can assume that
these are the first r + r' square brackets.
Then we obtain

Rv = £; Ak Ofk
r
o~v + £; Aklkv
r',
(v = 1, ... ,r + r'). (6.1.12)
270 6. The fundamental equation of dynamics. Analytical statics

By virtue of the general assumptions about the constraints, these equa-


tions can be resolved for Ak and A~.
Equality (11) now takes the form

(6.1.13)

with 3N -r-r' variations being independent of each other. For this reason,
the coefficient of each of these variations in eq. (13) must vanish which yields

(6.1.14)

The combination of eqs. (12) and (14) proves the above suggestion.
In summary, if the elementary work of all the constraint forces due to
virtual displacement of the system particles is equal to zero, then the 3N
constraint forces are expressed in terms of r + r' constraint multipliers Ak
and A~ which results in the following equations of motion for the particles
of system (3)

(v = 1, ... ,3N) (6.1.15)

or in vectorial form
r r'
miwi = mii\ = Fi + L Ak gradi fk + L A~eki' (6.1.16)
k=1 k=1

These equations are called Lagrange's equations of the first kind. The prob-
lem is determinate since the 3N equations (15), along with r+r' constraint
equations (4) and (5), have the same number of unknown variables

In what follows we will return repeatedly to the theorem proved here:


the sufficient condition for the following equation

(6.1.17)

with the r linear dependences between the variations


n
Laksbxs =0 (k = 1, ... ,r) (6.1.18)
s=1
6.2 Ideal constraints 271

is that functions Xs are represented in terms of the r independent La-


grange's multipliers AI, ... , Ar
r
Xs = 2: Akaks (8 = 1, ... ,n). (6.1.19)
k=l

Conversely, if eq. (17) holds, one can always find the r coefficients Ak
that represent functions Xs by linear relationships (19).

6.2 Ideal constraints


Holonomic and non-holonomic constraints satisfying the condition under
which the elementary work of the constraint forces for any virtual displace-
ment of the system particles vanishes, are referred to as the ideal constraints
or constraints without friction.
The constraints due to the contact of two smooth surfaces, the constraints
ensuring a constant distance between the system particles etc. are ideal.
For example, the internal forces of interaction of particles of a rigid body
are the constraint forces. By virtue of the action and reaction law the
resultant force and the resultant moment of the internal forces are equal
to zero and thus their elementary work due to eq. (5.2.5) is zero, too. This
means that the constraints in any rigid body are ideal. Reaction R of a
smooth surface, no matter whether fixed or moving, on a body moving on it
is directed along the normal to the surface and therefore is perpendicular
to the virtual displacement Dr of the point of contact of the body and
the surface. The elementary work R . Dr is zero and the smooth surface
presents an ideal constraint. An absolutely rough surface on which a body
rolls without slipping is an example of an ideal non-holonomic constraint.
In this case R· Dr = 0, too, since the second multiplier which is the virtual
displacement of the point of contact of the rolling body with the rough
surface is equal to zero.
From D'Alembert and Lagrange, the dynamics of a constrained system
of particles is based on the assumption that the constraints are ideal. The
first reason for this is that the achieved accuracy is sufficient to describe
the natural phenomena and motion of the technical systems. Secondly,
this assumption allows one to remain in the framework of basic principles
of Newton and D'Alembert and create the theory of motion of material
objects dealing only with the active (prescribed) forces.
For instance, we can take into account the non-ideal constraints at con-
tacts, for example the friction forces, by considering these forces as active
forces. This is caused by the necessity of using an experimental law of fric-
tion as this law eliminates inconsistencies in the number of equations of
mechanics and the number of unknown variables in the case of non-ideal
constraints.
272 6. The fundamental equation of dynamics. Analytical statics

In what follows, unless the other is stated, we adopt the assumption of


ideal constraints, that is the condition under which the sum of elementary
works of the constraint forces vanishes for any virtual displacement
3N N
LRvDC = LRi' Dri = O. (6.2.1)
v=1 i=1
Replacing here 8ri by means of eq. (1.6.7) or eq. (1.6.14) in terms ofvari-
ations of the generalised coordinates or variations of the quasi-coordinates,
respectively, we obtain

n N a
LDITs LRi' a ri = o. (6.2.2)
s=1 i=1 IT s
By definition (5.1.3) and (5.1.7) the internal sums in these equalities are
the generalised forces for the constraint forces
N N
Q* = """ R .. ari Ps* L R i'a=O,
""" ari (6.2.3)
s Li=1 " aqs ' =
i=1 IT s

and eq. (2) takes the form


N N
LQ;Dqs = 0, LP;DITs = O. (6.2.4)
s=1 s=1
In particular, if q1, ... , qn are independent generalised coordinates and the
non-holonomic constraints are absent, then variations Dqs as well as varia-
tions of the quasi-coordinates 8IT s are independent. It follows from eq. (4)
that all of the generalised constraint forces are equal to zero

(6.2.5)

6.3 The fundamental equation of dynamics and


Lagrange's central equation
We proceed from the differential equations of motion for the system of
particles (1.1)

(6.3.1)

The fundamental equation of dynamics is derived by means of eliminating


the constraint forces from the above equation. The problem of determining
the motion subjected to active forces is thus separated from the problem
6.3 The fundamental equation of dynamics and Lagrange's central equation 273

of determining the constraint forces, at least in the case of holonomic con-


straints.
Elimination of the constraint forces is achieved with ease in the case of
ideal constraints. Recalling definition (2.1) ofthe ideal constraints it suffices
to multiply each equation in (1) by 8ri and sum up the results, to get
N N
(6.3.2)
i=l i=l

Lagrange referred to this equation as the fundamental equation of dynam-


ics. Its derivation was based on the principle of constraint release enabling
construction of eq. (1) and on the definition of the ideal constraint. No
restriction on the kinematic properties of the constraints was imposed. For
this reason the fundamental equation of dynamics is applicable both to
holonomic and to non-holonomic ideal constraints.
In accordance with eq. (1.4.5) in the case of m redundant coordinates we
can write

(6.3.3)

and the fundamental equation reduces to the form

Introducing the generalised force we have

(6.3.4)

Variations 8qs are related by m + r' equalities


n+rn8Fk
2:a8qs=O (k=l, ... ,m), (6.3.5)
s=l qs

n+rn
2: a ks 8qs = 0 (k = 1, ... ,r'), (6.3.6)
s=l
which are obtained from m finite and r' non-holonomic conditions (1.4.8).
Using the theorem of Sec. 6.1 we arrive at the system of n + m equations

(s= 1, ... ,n+m),

(6.3.7)
274 6. The fundamental equation of dynamics. Analytical statics

containing n + 2m + r' unknown variables, among them n + m generalised


coordinates, m redundant coordinates and r' non-holonomic constraints.
The number of variables corresponds to the number of equations when
m + r' equations (1.4.8) are appended.
Let us recall that in eqs. (4) and (7) Qs designate the generalised forces
corresponding to the generalised coordinates qI, ... , qn+m, with m being
redundant.
In the case of holonomic system and absence of redundant coordinates
eq. (7) simplifies and takes the form

(s=l, ... ,n). (6.3.8)

The sum

"fAk 8 Fk
k=l 8qs

in eq. (7) can be omitted even in the case of redundant coordinates among
the generalised coordinates as this sum can be included into the following
sum

With this in view we can denote the number of the generalised coordinates
by n regardless of the fact that they are independent or related by finite
equations.
We proceed now to another form of the fundamental equation of dynam-
ics known as Lagrange's central equation.
We have

m{Vi' 8ri = !mivi' bri - mivi' (8rit

!mivi' 8ri - mivi' 8Vi + miVi' [8Vi - (8rit] .

The last term vanishes if the operations of varying and differentiation are
interchangeable. Noticing that
1 1 2
V· . 8v· = -8 (v· . v·) = -8v·
• • 2 • • 2"
we reduce the fundamental equation of dynamics (2) to the form

d N
-dt "m·v·
~ ••
i=l
. br'

=
1 N
8- "m·v 2
2~ "
i=l
+" N
~.
i=l
p. ·8r·

6.4 Rearrangement of Lagrange's central equation 275

or

d N ,
-dt "
~
mv ·
t t
·l5rt = I5T + 15 W , (6.3.10)
i=1

where T and I5T denote respectively the kinetic energy and its variation and
I5'W is the elementary work of the active forces. This equation is termed
as Lagrange '8 central equation.
When the operations of varying and differentiation are not interchange-
able, i.e. the law (1.7.5) does not hold, we obtain the following equation

d N N
-dt~
"m·v
t ·t ·l5rt = I5T + I5'W +"
~
m·v
t t . [(l5r)-
t -l5v·]
1., (6.3.11)
i=1 i=1

which is, following Hamel, referred to as the fundamental central equation.

6.4 Rearrangement of Lagrange's central equation


Consistent with eq. (5.1.1) the following expression
N

L mivi · 15ri (6.4.1)


i=1

can be treated as the elementary work of the linear momenta mivi (also
known as impulses) due to the virtual displacements 15ri of the system par-
ticles. Similar to the generalised forces Qi we can introduce the generalised
momenta. They are expressed in terms of the momenta mivi in a manner
like the generalised forces are expressed in terms offorces F i , i.e. by means
of eq. (5.1.3)

(6.4.2)

The equality
N n
L mivi . 15ri = LP l5qss (6.4.3)
i=1 s=1

corresponds then to relationship (5.1.4). Using transformations (1.3.5) we


can recast eq. (2) in the form

(6.4.4)
276 6. The fundamental equation of dynamics. Analytical statics

or
aT
Ps = aqs (s=l, ... ,n). (6.4.5)

The generalised momenta is thus equal to the derivative of the kinetic


energy with respect to the generalised velocity as formula (4.2.1) suggests.
The Lagrange's central equation (3.10) reduces to the form

(6.4.6)

When the forces are potential then due to eq. (5.3.9)


8T + 8'W = 8T - 8II = 8 (T - II). (6.4.7)
The function of the generalised coordinates and time equal to the differ-
ence of the kinetic and potential energies is named the kinetic potential or
Lagrange's function. It is denoted by
L (ql, ... ,qn, ql, ... ,qn; t) = T - II. (6.4.8)
Lagrange's function can contain time t as both kinetic energy and the
generalised potential energy can depend on time explicitly.
Thus, in the case of potential forces Lagrange's central equation is put
in the form

(6.4.9)

When the rule "d8 = 8d" does not hold the right hand sides of eqs. (6) or
(9) should be completed by the term
N N n
L miVi . [(8ri r - 8Vi] = LL ~ri [(8qsr - 8qs]
mivi .
i=1 s=1 qs
i=1
n Nan
= L [(8qsr - 8qs] L mivi· ari = LPs [(8qsr - 8qs]. (6.4.10)
s=1 i=1 qs s=1
Here formulae (1.7.3) and (2) are used. The central fundamental equation
is then written as follows
d n n n
dt LPs8qs = 8T + L Q s8qs + LPs [(8qsr - 8qs] . (6.4.11)
s=1 s=1 s=1
This equation should be expressed in terms quasi-velocities and varia-
tions of the quasi-coordinates. Repeating transformation (5.1.6) yields
N n n

L m i v i· 8r i = LPs8qs = LP;87rs, (6.4.12)


i=1 s=1 s=1
6.4 Rearrangement of Lagrange's central equation 277

where P: is related to Pr by means of the equality which is analogous to


(5.1.8)
n n aT aT
P: = 2: b8rPr = 2:bsr{F" = -a (8 = 1, ... ,n). (6.4.13)
r=l r=l qr Ws
The quantities P:
which are the derivatives of the kinetic energy with re-
spect to the quasi-velocities are referred to as the generalised momenta
corresponding to the quasi-velocities. For instance, in the case of a rigid
body having a fixed point
aT 0
~ = 8 U Wl + 8 l2 W 2 + 8 l3 W3 = Kl ,

a = 8 21 Wl + 8 22W 2 + 8 23W 3 = Kfj, (6.4.14)


a"lj 0
-a
W3
= 8 31 W 1 + 8 32 W 2 + 8 33 W 3 = K3 ,

as follows from expressions (4.7.4) and (4.8.13). Therefore, the momenta


corresponding to the projections of the angular velocity are the projec-
tions of the resultant angular momentum about this fixed point on the
corresponding axes. This illustrates the importance of the quantities in P:
mechanics.
The inverse relationship to eq. (13) is
n
P8 = 2:ar8P; (8 = 1, ... ,n). (6.4.15)
r=1

Thus making use of formulae (1.8.4) we have


n n n
2:Ps [(8qsr - 8q8] = 2:P; ars [(8qsr - 8qs] 2:
s=l r=l s=l

= ~P; {(87frr - 8wr - t t -Y;IWt 87f1 - te[87f1} ' (6.4.16)

and the central equation (11) takes the form


dn{)T n naT_
dt ~ aws 87fs 8T + ~Ps87fs + ~ aws [( 87f s) - 8ws] (6.4.17)
nnn aT nn aT
- ~ ~ ~ -Y;s aW r Wt 87f s - ~ ~ e~ aW r 87f s·
If the active forces are potential forces, the right hand side can be cast as
follows
n
8T + 2: Ps87f s = 8 (T - II) = 8L. (6.4.18)
s=1
278 6. The fundamental equation of dynamics. Analytical statics

Finally, as the potential energy does not depend upon the generalised
velocities, expressions (5) and (13) can also be written in the form

* aT aL (6.4.19)
p =-=-
8 aW 8 aw 8 ·

6.5 Equilibrium of the system of particles


When the system is in equilibrium, the acceleration Wi of any particle of
the material system with respect to the inertial axes Oxyz is equal to zero.
Applying the principle of constraint release we obtain instead of (1.1)

(6.5.1)

Using eq. (1.15) we can write 3N equations of equilibrium

(6.5.2)

Of course, in assuming an equilibrium we assume that the constraint equa-


tions (1.4) and (1.5) do not contain time explicitly and moreover the free
terms 9k are absent in eq. (1.5) otherwise all the velocities can not equal
zero simultaneously.
Given prescribed forces, we have 3N equations of equilibrium with 3N +
r + r' unknown variables which are the 3N Cartesian coordinates of the
particles and r + r' constraint multipliers. In addition to the equations
at our disposal, we have r equations of the finite constraints. Equations
for the non-holonomic constraints (1.5) are satisfied identically in equilib-
rium when all ~v vanish. The problem of equilibrium of the system subject
to ideal constraints is determinate only in the case of no non-holonomic
constraints. In other words, the problem is no longer determinate in the
presence of such constraints as r' of the 3N + r + r' quantities remain
indeterminate.
Describing the system position by n + m generalised coordinates related
by m finite equations of constraints

(6.5.3)

and r' equations of the non-holonomic constraints

n+rn
L ak8q8 = 0 (k = 1, ... ,r'), (6.5.4)
8=1
6.5 Equilibrium of the system of particles 279

we obtain by setting in the left hand side of eq. (3.7) Wi = 0, the n +m


equations of equilibrium

(6.5.5)

These express the condition that the sum of the prescribed generalised
forces and the generalised constraint forces

(6.5.6)

vanishes. The problem is defined when non-holonomic constraints are ab-


sent. The first sum on the right hand side of eq. (6) can be written in the
form

(6.5.7)

since, by virtue of eq. (3), the following sum

can be cancelled out.


Let us consider the case of potential forces, then

Qs __
- OII (
s=1 , ... ,n+m.
)
oqs

with II denoting the potential energy of the system. Accounting for eq. (7)
we can write the equilibrium equations in the case of no non-holonomic
constraints in the form

00 (-II +
qs
f
k=l
)..kFk) = 0 (s = 1, ... , n). (6.5.9)

These are extremum conditions of the function

-II (q!, ... , qn+m)

subject to m constraints (3).


In the case of no redundant coordinates and non-holonomic constraints,
the equilibrium equations

Qs=O (s=l, ... ,n) (6.5.10)


280 6. The fundamental equation of dynamics. Analytical statics

express the conditions under which the generalised forces vanish. When
they are potential forces, the latter equation becomes

all =0 (s = 1, ... ,n ) , (6.5.11)


aqs
determining stationary values of the potential energy of the system. In
the equilibrium position of the system under the ideal constraints and the
potential forces only the potential energy of the system takes a stationary
value. In other words, when the system is displaced from the equilibrium
position q~, ... , q~ into an infinitesimally close position

q~ +8ql,'" ,q~ +8qn


the increment in the potential energy
~II = II (q~ + 8ql, ... , q~ + 8qn) - II (q~ , ... , q~) (6.5.12)
is the value of the second order or higher terms in variations qs of the
generalised coordinates. The first order terms in the expansion of ~II in
terms of qs vanish by virtue of eq. (11).
The character of the equilibrium of the system subjected to the potential
forces is determined by the character of the extremum of function II. The
equilibrium can be stable or unstable. Due to the fundamental theorem of
Lagrange and Dirichlet, the equilibrium position is stable if the potential
energy in the equilibrium position possesses a minimum. The inverse state-
ment concerning the case in which the extremum is not a minimum was
proved by Lyapunov and Chetaev. The presentation of this theory and the
rigorous definition of the equilibrium are beyond the scope of this book. It
is the subject of the special treatise on the theory of stability of motion,
e.g. [20] and [62].
We formulated here the problem of equilibrium by using the principle
of constraint release under the condition of zero acceleration (Wi = 0) of
the system particles and referring to the consequences of the fundamental
equations of dynamics (3.2). Setting Wi = 0 in this equation leads to the
principle of virtual work
N

l:Fi ·8r i=0 (6.5.13)


i=l

which states that the sum of work done by the active forces due to virtual
displacements of the system particles from the equilibrium position is zero
provided that the system is subject to ideal constraints.
If some of the forces are potential forces, the principle of virtual work
can be set in the following form
N
-8IT + l : Fi · 8ri = O. (6.5.14)
i=l
6.6 Examples of deriving equilibrium equations and constraint forces 281

Let us consider eq. (6) when the non-holonomic constraints are absent.
We choose the generalised coordinate so that the constraint equations be-
come as simple as possible

FI = qn+1 - q~+1 = 0, ... ,Fm = qn+m - q~+m = 0,


where q~+k are constant values of the redundant coordinates. Then

8F k
8qs -
_ ° _{O,
n+k,s - 1,
n + k =I- s,
n + k = s,
and the equilibrium equations take the form
m
Qs + LOs,n+kAk = 0
k=1

or, more specifically,

QI = 0, ... , Qn = 0, Al = -Qn+I, ... , Am = -Qn+m. (6.5.16)


For example, let us consider the equilibrium of a free particle on a sur-
face. Entering the curvilinear coordinates ql, q2, q3, such that q3 = q3 cor-
responds to the surface in question, we calculate the elementary work of
the active force acting on the particle. Using notation (B.7.2) we obtain

O'W = F· Or = F· rsoqS,
where r s denotes the base vectors of the surface. The equilibrium equation
(14) yields
F· rl = 0, F· r2 = 0, Al = -F· r3.
These equalities hold on the surface, i.e. at q3 = q3. Expressing the vector
of the active force in terms of its covariant components
F =Fsrs = F . r srs = F . r3r3
we see that the equilibrium is feasible only if the active force is directed
along the normal to r3 to the surface. As F = - R we conclude that the
constraint multiplier is equal to the covariant component R3 of the reaction
force.

6.6 Examples of deriving equilibrium equations


and constraint forces
6.6.1 System of three rods
Let us consider a system of three rods OA, AB, Be attached to each other
by joints A and B, see Fig. 6.1. The joint 0 is fixed and the system is held
282 6. The fundamental equation of dynamics. Analytical statics

FIGURE 6.1.

in the vertical plane Oxy by means of three threads, the angles between the
rods and the downward vertical being <P~, <P~, <P~, respectively. The thread
tensions T l , T 2 , T3 are required.
The active forces are the rod weights G l , G 2 , G 3 . The potential energy
of the system is

where Xl , X2 , X3 denote the coordinates of the centre of gravity of the re-


spective rod

Xl 81 cos <PI' X2 = 11 cos <PI + 82 cos <P2,


X3 h cos <PI + 12 cos <P2 + 83 cos <P3·

Here h, h, b denote the rod lengths and 81, 82, 83 the distances between
their centres of gravity and joints 0, A , B, respectively. The angles between
the rods and the downward vertical under mental release of the constraints
due to the threads are denoted by <PI' <P2' <P3. These three constraint equa-
tions can be cast in the form of eq. (5.15)

We obtain then that


6.6 Examples of deriving equilibrium equations and constraint forces 283

and by virtue of eq. (5.16) we obtain the constraint multipliers

Al = (G181 +G2 h +G3h)sin<p~, }


A2 = (G282 + G 312) sin<p~, (6.6.1)
A3 = G383 sin<pg.

The constraint multipliers are the generalised constraint forces but not the
required thread tensions. In order to find the latter, we notice that the
elementary work of the generalised constraint forces coincides with that
done by the thread tensions due to virtual displacements of points A, B, C
(where these forces are applied) from the equilibrium position

(6.6.2)

Since

DYA h cos <PID<PI,


DYB = h cos <PID<PI + 12 cos <P2D<P2,
Dye h cos <PID<PI + 12 cos <P2D<P2 + b cos <P3D<P3
we equate the coefficients in eq. (2) for the independent variations D<ps, to
obtain a system of three equations

where ips was replaced by <p~. Solving this system for Ts yields

TI = ~GI ;~ + G2 + G3) tan<p~ - (G2 ;: + G3) tan<p~,


82 ) 83 (6.6.3)
G2Z; + G3 tan <P2 - G3l; tan <P3,
0 0
T2 =
83 0
T3 = G3l; tan<P3'

It is easy to prove that the system is in equilibrium since the resultant


moment of the gravity forces G s and the thread tensions Ts about the
fixed joint 0 is zero.

6.6.2 Equilibrium of a heavy rod gliding by their ends on a


smooth surface
Rod G of length 1 can glide by its ends A and B on the internal surface
of a cup, Fig. 6.2. Directing axis Oz along the upward vertical we describe
this surface by the following equation

z = f (x, y). (6.6.4)


284 6. The fundamental equation of dynamics. Analytical statics

FIGURE 6.2.

The equilibrium position of the rod and the reactions forces of the walls of
the cup are required.
Denoting the coordinates of the rod ends by Xl, YI, Zl and X2, Y2, Z2 we
can write the three constraint equations

(6.6.5)

Let P and q designate the partial derivatives offunction f(x, y) with respect
to X and y, respectively. The subscripts 1 and 2 indicate that these values
are referred to points A and B, respectively. Consistent with Sec. 1.4 it is
necessary to consider the matrix

-PI -ql 1 0 0 0
0 0 0 -P2 -q2 1
Xl - X2 YI - Y2 Zl - Z2 X2 - Xl Y2 -YI Z2 - Zl

The determinant of the following 3 x 3 submatrix


-ql 1 0
o 0 -P2
YI - Y2 Zl - Z2 X2 - Xl

is not zero as will be shown below. Thus, equations (5) can be resolved for
three of the six introduced quantities.
Let us denote the unit vector of the inward normal to the surface (4)
as ill, see Fig. 6.2, and the directions cosines of the normal relative to the
coordinate axes as a, (3, "(. Then
P (3 = - q
a=
VI +p2 +q2 '
6.6 Examples of deriving equilibrium equations and constraint forces 285

Let e denote the unit vector directed along the rod from A to Band
ell e2, e3 the cosines of the angles between e and the coordinate axes
Z2 - Zl
e3 = --Z-· (6.6.7)

The centre of gravity of the rod is assumed to be at the mid-point of the


rod, then the potential energy of the gravity force is given by
1
II = "2G (Zl + Z2) .
Using the notation of eqs. (6) and (7), the equilibrium equations (5.9)
are
al>\! = 2A3Zen1,
/3lAl = 2A3Ze21'1, (6.6.8)
1
Al = 2A3Ze3 + "2G,
Resolving them for AI, A2, A3 yields

.!.G al =.!.G /31


2 el 'Y1 - e3a 1 2 e2'Y1 - e3/3l
-.!.G a2 - -.!.G /3 2 (6.6.9)
2 el 'Y2 - e3 a 2 - 2 e1 'Y2 - e3/32 '

(6.6.10)

el a1 a2
(6.6.11)
e2 /31 /32·

This relationship also satisfies two of the equalities (9). It expresses the
fact that the normal vectors m1 and m2 lie in the vertical plane passing
through the rod. Indeed, denoting the unit vectors of the upward vertical
and the normal vector to the above plane as k and n, respectively, it is
sufficient to prove that n is perpendicular to m1 and m2. We have
exk
n=---
Ie x kl
and thus
286 6. The fundamental equation of dynamics. Analytical statics

which completes the proof.


The denominators in eqs. (9) and (10) are proportional to the projections
of vector n on axes Ox and Oy. Indeed, due to the above

and thus

n - en· - e3i3'
~ ~

I - IeXIDi I'
Vector n is parallel to the plane Oxy which means that at least one of these
expressions is non-zero. We notice in passing that the determinant of the
above submatrix is proportional to nl.
Another relationship which is a consequence of eq. (9) is
'"Y1 + "Y2 =2e3 or (6.6.12)
ctl ct2 el

Its meaning will be explained in what follows.


Therefore, we obtained three equations, eqs. (11) and (12), and the equi-
librium positions exist provided that these equations along with the con-
straint equations (5), have solutions.
Now we proceed to determine the constraint forces. The resultant of Ri
of the constraint forces at the ends A and B of the rods is determined with
the help of eq. (1.9) which yields

RI = Al grad l FI + A3 grad l F3, R2 = A2 grad 2 F2 + A3 grad2 F3,


as FI and F2 depend only on xI, yI, Zl and X2, Y2, Z2, respectively. The first
terms in these expressions describe the reaction forces NI and N2 of the
cup, while the second terms describe the rod reactions TI = -T 2 • We
obtain for i = 1,2

and

(6.6.14)
Taking into account conditions (11) and (12) it is easy to prove that the
forces N I, N 2 and G are in equilibrium.

6.6.3 Rod in an elliptic cup


This is a particular case in which the cup surface is an ellipsoid of revolution
about axis Oz. This problem known as Brashman's problem is studied in
[101].
6.6 Examples of deriving equilibrium equations and constraint forces 287

Placing the origin of the coordinate system as the ellipsoid vertex we


have

./ X2+y2
z= c+cy 1- a2

Dealing with the surface of revolution one can take the plane Ozx as the
plane in which the rod lies. It suffices to put YI = Y2 = in the above
formulae. The equations for determining the unknown variables Xl and X2
°
are

where

~
z = c+cy1- ~.

Since

1. = _~ =
a p
a2
ex
J 1_ X2
a2

we put X = a sin U in the above equations, to have

.
( smU2 .)2
- smUl + 2ac2 (COSU2 - COS U I
)2
= [2

a
Introducing the half-sum and the half-difference of the angles UI and U2

f3 = U2 - UI
2 '
we can rearrange the latter equations

sin a (cos 2 f3 - e2 cos2 a) = 0,


where

°
Consider an elongated ellipsoid of revolution, i.e. c > a, < e2 < 1.
The equilibrium position a = is not feasible as a > 0, thus the feasible
°
equilibria are given by a = 7r and UI + U2 = 27r. As Fig. 6.3 shows, it
288 6. The fundamental equation of dynamics. Analytical statics

FIGURE 6.3.

corresponds to the horizontal equilibrium position AIBI in which angle {3


is defined by the following equation
Z2
sin 2 {3 (1 - e2 ) = -2
4a
and the problem has a solution (which implies that an equilibrium position
exists) when the obvious condition

lS;2a~
holds.
The second equilibrium position A2B2 takes place when cos 2 {3 = e2 cos 2 a,
which is equivalent to l = 2a sin 2 (3. It can be proved that the rod passes
through the focal point of the ellipse. A simple geometrical proof is sug-
gested in [101).
If a flat ellipsoid (c > a, 0 < e2 < 1) is considered only the horizontal
equilibrium (a = 7r) is feasible.

6.6.4 Equilibrium of a rigid body in a central force field


The potential energy of a body is given by expression (5.5.6) where r =
const as we are considering an equilibrium. The aim of the forthcoming
analysis is to find under which values of the parameters defining the ori-
entation of the trihedron of central axes Gxyz (see Fig. 5.4a) relative to
axes of fixed directions M~'f/( , the potential energy takes stationary values.
These parameters are Euler's angles {), 'ljJ, r.p which appear in the potential
energy in terms of x / r , y/r, z / r coinciding with the direction cosines of the
angles between the vector MG and the axes M~'f/( .
6.6 Examples of deriving equilibrium equations and constraint forces 289

According to eq. (5.11) the required positions of trihedron Gxyz are


determined by the conditions
arr _ arr arr
mN =- a1} = 0, m3 =- a'IjJ = 0, m3 = - acp = o.

By means of formulae (5.2.18) this can be cast as follows

m1 cos cp - m2 sin cp 0,
(m1 sin cp + m2 cos cp) + m3 cos 1} 0,
m3 0,

where ml, m3, m3 are determined by eq. (5.5.10). Thus there exist two
possibilities

a) m1 = 0, m2 = 0, m3=0;
b) m3 = 0, sin1} = 0, m1 coscp - m2 sincp = O.
Case a) occurs under the conditions

(3'Y = 0, 'Ya = 0, a(3 = 0,

implying that two of the three direction cosines must vanish. This means
that in the equilibrium position one principal axis of inertia must be di-
rected along MG. For instance, if a = f3 = 0, 1"11 = 1, then x = y = 0 and
the only non-zero coordinate of point M, referring to the axes Gxyz, is z.
In this case axis Gz of the inertia ellipsoid is directed along MG.
In case b) we direct axis M( along MG, then ea = 'TJa = 0 in eq. (5.5.11).
As sin 1} = 0 we obtain

a = (3 = 0, 'Y = 1,

that is we return to one of the cases considered earlier.


The same equilibrium conditions can be obtained without calculating
the generalised forces but by means of the extrema of the potential energy.
According to eq. (5.5.6) and notation (4.6.8) the potential energy differs
from the quadratic form

F = 2
1 [ (2 - 6"1 - 6"2) a 2 + (26"1 - 6"2 - 1) (3 2 + (26"2 - 6"1 - 1)"(2] (6.6.15)

only in an additive constant and an immaterial positive factor. In accor-


dance with eq. (4.6.6) the small, middle and large principal axes of inertia
are directed along Gx, Gy, Gz, respectively.
By virtue of Sec. 4.6
290 6. The fundamental equation of dynamics. Analytical statics

FIGURE 6.4.

The coefficient in front of (32 can change its sign. The line L

shown in Fig. 6.4 splits the plane of values El, E2 into two domains. In the
domain under line L the above coefficient is positive, whereas above the
line it is negative.
The stationary values of the quadratic form F subject to constraint
q> = ex2 + (32 + "(2 - 1 = °
is sought from the equations
() () ()
()ex(F-Aq» =0, {)(3 (F - Aq» = 0, -(F-,\cI» =0
{)"(
which are recast as follows

(2 - El - E2 - A) ex = 0, (2El - E2 - 1 - A) (3 = 0, (2E2 - El - 1 - A) "( = 0.


(6.6.16)
Since ex, (3, "( can not be zero simultaneously there are three cases
a) ex = (3 = 0, A = 2E2 - El - 1 < 0,
b) (3 = "( = 0, A= 2 - El - E2 > 0,
c) "( = ex = 0, A = 2El - E2 - 1 ~ o.
These determine the equilibrium positions of the body. It is also known that
the obtained values of A are equal to the extremum values of F. Indeed,
multiplying the equalities in (16) by ex, (3, ,,(, respectively, and summing up
the products yields

Therefore, case a) corresponds to the minimum value of the potential en-


ergy and case b) to the maximum value of the potential energy. By virtue
6.6 Examples of deriving equilibrium equations and constraint forces 291

of Lagrange-Dirichlet's theorem case a) describes a stable equilibrium po-


sition in which the "long" axis of the inertia ellipsoid lies along MG. Case
b) corresponds to the coincidence of the directions of the" short" axis and
MG and, due to Lyapunov' theorem, this equilibrium position is unstable.
The character of the equilibrium in case c) depends on the relationship
between the values of the moments of inertia.

6.6.5 Equilibrium of a rigid body suspended on elastic rods


A rigid body is suspended on a system of elastic weightless rods attached
to the body by means of spherical joints as points M1, ... , Mm. The other
ends of the rods are fixed at immovable points 81, ... , 8 m . It is necessary to
find how the body is displaced from the initial positions under loading by
the forces increasing monotonically and slowly from zero up to the values
F 1, ... , F n. In other words, the transition from the initial unloaded state into
the final state is presumed to occur as a continuous sequence of equilibrium
configurations. The rods are not prestressed.
Let Oxyz denote the system of axes fixed in the body, V and rn o denote
the resultant force and the resultant moment about the pole 0 of the
system of forces F l , ... , F n, respectively, and rk denote the position vectors
OM~ of joints M k . A similar problem for a system of rods with a single
common joint is considered in Sec. 5.8. In contrast to the previous problem
we should now take into account that the displacement vectors Uk of the
joints Mk are different and are given by

Uk = Uo + (J x rk.
We restrict our consideration to small rotations and displacements. Then
the elongation of the k - th rod up to the first order of these values is

(6.6.17)

and thus

if = Uo . e2e2 . Uo + (J . (e2 x rk) (e2 x rk) . (J - 2uo . e2 (e2 x rk) . (J.


The potential energy of the elastic rods can be written in the form

II = (6.6.18)

This is a quadratic form of the projections of the vectors Uo and (J. Within
the accuracy assumed it makes no difference whether the axes are fixed in
the body or in space. The axial rigidities of the rods are designated by Ck.
292 6. The fundamental equation of dynamics. Analytical statics

Being the sum of positive components ~ckR, the potential energy is


a positive definite form. Sylvester's criterion always holds true. For this
reason, the equilibrium position, due to the Lagrange-Dirichlet theorem, is
stable as long as the forces F 8 are absent.
Due to eq. (5.14), the following equation is valid

-8II + V . 8uo + rno ·88 = 0

at equilibrium. We thus obtain

These equilibrium equations can be directly constructed without calculat-


ing the potential energy. To this end, we write down the condition un-
der which the resultant vector and the resultant moment of the forces
- T 1, •.. , - T m, F 1, ... , F n about pole 0 vanish. It is sufficient to notice that
the constraint force -Tk of the rod and its moment about pole 0 are

and to replace /k by its expression from eq. (17).


Equalities (19) can be written in matrix form as

where A, B, C denote the following matrices


m

A= LCkegeg,
k=l

Determination of the column-matrices Uo and () is thus reduced to solving


a system of six linear equations or calculation of the inverse of matrix Q.
The inverse matrix exists since the determinant IQI is not zero. Due to
Sylvester's criterion it is positive.

6.6.6 A special case of a prestressed system


We consider the same body as above, i.e. the body suspended on elastic rods
and at equilibrium under the action of some prescribed forces and the con-
straint forces, [71]. The resultant vector and the resultant moment about
the pole 0 are denoted by Va and rnf?, respectively. The rod elongations
6.6 Examples of deriving equilibrium equations and constraint forces 293

FIGURE 6.5.

from their initial states are 8k . We consider passage from the initial equilib-
rium configuration 8 0 to configuration 8 due to a new resultant vector V
and the resultant moment rn o about the same pole O. In other words, we
look for the displacement vector Uo of the pole 0 and the rotation vector
(J describing this passage.
As shown in Sec. 5.8 the potential energy of the k-th rod in configuration
8 is as follows

Thus, the elongation should be found up to the values of second order.


Taking the rotation vector (J to be small, it is necessary to retain in Ro-
drigues's formula (3.1.11), for displacement of point M2,
only the terms of
second order in (J. Then we obtain

Figure 6.5 shows that

The solution of this equation is sought in the form

in which the first term coincides with that in eq. (17) and the second term
is the correction term of second order. This substitution yields
294 6. The fundamental equation of dynamics. Analytical statics

An expression for 8kfk calculated within the above accuracy is

8k!k = 8k [uo· e~ + O· (rk x e~)] + :zg {Iuo x e~12 + le~ x (0 x rk)1 2


-2 (uo x e~) . [e~ x (0 x rk)] + zg (eg x 0) . (0 X rk)}. (6.6.20)
Let us adopt that the forces F~, ... F~ impressed on the initial configura-
tion retain their magnitudes and directions during the passage into the new
configuration S. Due to eq. (5.3.23) the change in the potential energy of
this system of forces calculated up to the squared values included is equal
to
o 1 0
~II2=-Vo·uo-mo .O+"20.Q ·0,

where tensor QO is given byeq. (5.3.24).


The first term in the expression for the potential energy of the rod system

is given by eq. (18) as the additional term in expression (2) for fk may give
only a correction term of the third order in the equation for R. The second
term is denoted by ~IIl. The third term represents an additive constant
which is of no importance and can be omitted since it is the value of the
potential energy in configuration So. We now have
m
~IIl + ~II2 = L ck8ke~ . (uo + 0 x rk) - (Vo· uo + m~ .0) +
k=l
~ --zo
"21 L.,.. Ck 8k [I uo x ek01 2 + Iek0 x (0 x rk) 12 -
k=l k

2 (uo x e~) . [e~ X (0 X rk)] + zg (eZ X 0) . (0 X rk) + ~O. QO. oJ.


(6.6.21)
As So is an equilibrium configuration the variation of the potential energy
at uo = 0 and 0 = 0 must be equal to zero. This variation is

(8II)uo=0,li=0 = (f>k 8ke Z - YO) .8uo + (fCk8krk x eZ - m~) ·80.


k=l k=l
In view of the arbitrariness of variations 8uo and 80, the condition under
which this variation vanishes, yields the following equation
m m

L ck8ke~ - Vo = 0, Lrk x ck8ke~ - m~ = o. (6.6.22)


k=l k=l
6.6 Examples of deriving equilibrium equations and constraint forces 295

These equalities express the fact that the constraint forces in the initial
configuration So equilibrate the applied forces.
The consequence of equalities (22) is that the linear terms in the expres-
sion for the potential energy vanish. Now using eqs. (18) and (21) we arrive
at the expression for the potential energy

(6.6.23)

where expressions for matrices A, B, C are given above and matrices AI,
B l , C l are as follows

(6.6.24)

Here rk and r~ denote the position vector of the joints mounted on the body
and of the points of application of forces F~ in the equilibrium configuration
So, e~ and l~ are the unit vector of the rod skMgand its length in this
configuration. rk denotes a skew-symmetric 3 x 3 matrix accompanying
vector rk due to the rule (A.2.3) and a and a' are respectively the column
matrix and the row-matrix of the projections of vector a.
The column-matrices of the projections of the vectors of displacement
and rotation should be found by analogy with the previous example by
replacing the corresponding matrices by A + AI, B + B l , C + C l .
The question of the sign of the quadratic form (23) which is
1 m
"2 L (ckff + 2bkik) + L).II 2
k=l

can not be solved in advance as it was done in the case of the system of
prestressed rods. The Sylvester criterion for the matrix

A+Al II
M= II B'+B~ B+Bl
C+Cl
(6.6.25)

can hold for certain initial elongations bk and the equilibrium configuration
So is stable. However the criterion can fail for other initial elongations and
the equilibrium configuration turns out to be unstable.
We restrict our attention to the case when the parameters satisfy all the
conditions of Sylvester's criterion except for one, namely that the determi-
nant IMI = o. Then the system of equations defining vectors Uo and () has
296 6. The fundamental equation of dynamics. Analytical statics

no solution for arbitrary V and rno and there are no equilibrium config-
urations 8 close to 8 0 . However it is known that a inhomogeneous system
of linear equations with zero determinant may have solutions under special
conditions imposed on the right hand sides of the equations. If one of the
first minor determinants is not zero then the solutions are determined up
to additive constants proportional to an arbitrary parameter c. Therefore
there exist such values of V and rno which give rise to a continuous series
of equilibrium configurations proportional to an arbitrary parameter. This
is what is referred to as an indifferent equilibrium.

6.6.7 Equilibrium in the presence of Coulomb '8 friction


A rigid plate is compressed by two plane surfaces on its faces. The action
line of an active force F applied to the plate lies in the mid-plane of the
plate. The limiting equilibrium of the plate under the frictional forces on
the faces is considered. This problem was studied by Zhukovsky in [100]
and by McMillan in [65]. The virtual displacement of the plate is given
by vector 8ro of the virtual displacement of the pole 0 of the axes Oxyz
and by vector (J = Bzi3 of the infinitesimal rotation about axis Oz. The
elementary work of the active force F in this virtual displacement is

8'W1 = F· 8ro + m~Bz,


where m~ denotes the moment of force F about axis Oz.
The value of the frictional force on the elementary surface do of the con-
tact of the plate and the plane surface is equal to jpdo, j and p denoting
the friction coefficient and the pressure, respectively. The direction opposes
the velocity v which the surface do would have when the limiting equilib-
rium is broken. The elementary work of the frictional forces due to two
contact surfaces 8 is then given by the following expression

8'W2 = -211 s
jp; . (8ro + hB z x r) do,

where r denotes the position vector of surface do. Introducing the position
vector r p of the instantaneous velocity centre P one can represent the
velocity vector v and its value v as follows

Denoting E = sign w z we have

8'W2 = - 2E8rO· 11
s
jph ~r ~ ;p~p) do - 2EBz 11 jpr·l~r_-r:l)
s
do.

According to the principle of virtual work


8'W1 + 8'W2 = O.
6.6 Examples of deriving equilibrium equations and constraint forces 297

Equating the coefficients of independent variations c5ro and () z to zero yields


two equations

o
m z = 2c
JrrJ jp r·(r-rp)
Ir _ rpl do, (6.6.26)
s
where the latter equality can be rearranged into the form

o
m z =2crp·
JrJr r - rp
jplr_rpldo+2c
JrJr jplr-rpldo. (6.6.27)
s s
Because

iarp· (r - rp) = rp x [h x (r - rp)] ,


we obtain, by means of eq. (26) for F,

2d3 rp· jrrJ jP 1rr-rp


_ rp1do = rp x F.
s
Let r' denote the position vector of a point on the action line of force F,
then

where m; is the moment of force F about the axis passing through the
instantaneous velocity centre parallel to Oz. Formula (27) can be recast in
the form

m; =2c 11
s
jplr-rpldo. (6.6.28)

Expressions for the projections of force F on coordinate axes is now needed.


As

it· [h x (r - rp)] = - (y - yp), i 2 • [h x (r - rp)] = x - xp

and

we obtain
298 6. The fundamental equation of dynamics. Analytical statics

Thus, introducing the function

W(xp,yp) = 211 fplr - rpl do = 211 fpV(x - xp) + (y - yp)2do,


s s
(6.6.30)

we can reexpress the force F and its moment mf as follows


Fy = -c~,
oW mf = cW(xp,yp). (6.6.31)
uxp

Assuming a distribution of normal pressure p(x, y) and calculating W we


can, by means of eq. (31) and coordinates of the instantaneous velocity
centre, find the value, the direction and the action line of force F which
breaks the equilibrium and causes an initial rotation about the instanta-
neous velocity centre. Given the value and the direction of force F, the first
and the second equalities in (31) yield Xp, yp and also the sign of W z . The
third equality in (31) enables one to find the action line of force F such
that the initial rotation occurs about the instantaneous velocity centre. We
notice also that the signs of mf and W z coincide as W > o.
Let us determine under what condition the initial displacement of the
plate will be pure translation. In this case the elementary work of the
frictional force is

8'W2 -2 11 fp :~ . (8ra + hBz x r) do


s
-2 11 fpdo :~ . 8ra - 2B z 11 fp (rdo x :~) . i3
S S

and the principle of virtual work yields

F = 211 fpdo :~ , m~ =211 fp(rdox :~) ·i3 . (6.6.32)


s s
If we introduce the centre of pressure, which is the point determined by
the following position vector

II fprdo
s (6.6.33)
re = II fpdo '
s
then the moment mCj can be written in the form

m~ = (re x F) . h. (6.6.34)
6.6 Examples of deriving equilibrium equations and constraint forces 299

Clearly, this moment is zero if the force passes through the centre of pres-
sure.
Therefore, an initial pure translation in the force direction takes place
when the. force reaches the maximum value

Fm = 211 s
jpdo, (6.6.35)

at which equilibrium is still possible and passes through the centre of pres-
sure.
Let us consider now the case of a pair of forces. Then F = 0 and due to
eq. (31)

{N! _ 0 a\It _ 0
axp - , ayp - , (6.6.36)

that is, \It has a stationary value. In order to determine the sign of the
second variation 82\It which coincides with the increment in \It when Xp, yp
are replaced by Xp + 8xp, yp + 8yp

2
8 \It = -2
1 (a 2\It
-a2 8xp
2+ 2 a a 2a\It 8xp8yp a
\It 2)
+ -a yp
2
28yp

11 ~;
xp xp yp

= ~ [(y - yp)2 8x~ - 2 (x - xp) (y - yp) 8x p 8yp+

11 ~;
s
(x - xp)2 8y~] do =~ [(y - yp) 8xp - (x - xp) 8yp]2 do.
s
The value in the square brackets can be equal to zero for any x and y
only if the elementary surface is the following straight line
y-yp
""---.......;;...- = const .
X-Xp

Excluding this case we obtain that the second variation 82\It > 0 and
the above stationary value is a minimum. Zhukovsky referred to the point
(xp, yp) defined by condition (36) as the frictional pole. When the body is
subject to the pair of forces

(6.6.37)

the initial rotation about the frictional pole occurs. No equilibrium is fea-
sible if the absolute value of the moment exceeds \It min. As function \It can
have only minima, the minimum obtained is the only one. For this reason,
only one frictional pole can exist. Let us notice in passing that Zhukovsky
[100] made a number of interesting suggestions about the properties of
equilibrium while studying the surfaces \It = const.
300 6. The fundamental equation of dynamics. Analytical statics

Under the uniform distribution of pressure over the contact surface, the
function W can be written in the form
w= 2fGr*, (6.6.38)
where G is the force pressing the plate and r* was introduced in Sec. 5.12
when the dissipation function of frictional forces was analysed. Calculation
of r* is performed by means of eqs. (5.12.12) and (5.12.13). In the case of
the contact over a circular plate of radius a

r* = 2fGa'IjJ
W = 2fGa-;; (R)
-;;: ,

where R = rp, expressions for 'IjJ for R :::; a and R ?: a are given by eqs.
(5.12.14) and (5.12.15), respectively, and the derivative 'IjJ' with respect to
argument Ria is given by eq. (5.12.20) in which the factor fG should be
omitted. The table of functions 'IjJ (Ria) and 'IjJ' (Ria) is shown below

I:
0 0.6667 0 1 1.1318 0.8488
0.1736 0.6817 0.1730 1.0154 1.1449 0.8550
0.3420 0.7247 0.3369 1.0642 1.1870 0.8714
0.5000 0.7897 0.4838 1.1547 1.2669 0.8942
0.6428 0.8677 0.6076 1.3054 1.4038 0.9199
0.7660 0.9487 0.7047 1.5557 1.6375 0.9452
0.8660 1.0227 0.7744 2 2.0631 0.9677
0.9397 1.0814 0.8188 2.9238 2.9665 0.9851
0.9848 1.1189 0.8420 5.7588 5.7797 0.9962
00 00 1

The table of functions 'IjJ ( : ) and 'IjJ' ( : )

Of course, the centre of the circle 0 is both the centre of pressure and
the frictional pole. The latter statement is due to 'IjJ' (0) = 0 and is also
clear from the symmetry of the problem. The initial translatory mQtion
takes place under a force having the value Fm = 2fG passing through
the centre of the plate whereas the initial rotation about the circle centre
occurs under the pair offorces with the moment Fma'IjJ (0) = 0.667Fma. In
the general case we can adopt that, due to the problem symmetry, the axis
Ox is parallel to vector F. By means of eq. (31) we obtain

xp ~O, e~ F~Fm~' (~),


1, m; ~aFm~ (~) ~a< ~~
6.6 Examples of deriving equilibrium equations and constraint forces 301

Since 0 < 'ljJ' < 1 the equilibrium is possible only if F < Fm. The action
line intersects axis Oy at the point

Zhukovsky also analysed the case of contact on two small areas.


7
Lagrange's differential equations

7.1 Derivation of Lagrange's equations of the


second kind
Differential equations of motion for the generalised coordinates can be ob-
tained easily with the help of Lagrange's central equation. The equations
will be derived twice here. The first derivation will assume that the oper-
ations d and 8 are not interchangeable, while the second one will assume
that the operations are interchangeable. In the first case, i.e. if d8 =I- 8d, it
is necessary to use Lagrange's central equation. Taking into account that
the kinetic energy is a function of the generalised coordinates and velocities
we can write

(7.1.1)

We also have

! t Ps8Qs = t
s=l s=l
[P s8qs + Ps (8Qst] , (7.1.2)
304 7. Lagrange's differential equations

and substitution into (6.4.11) yields

n
+L [Ps (bqs)" - PsbQs] .
8=1

The underlined terms cancel out and we arrive at the equality

(7.1.3)

Let us consider now the case in which db = bd. We now use Lagrange's
central equation in form (6.4.6) which, after repeating the above derivation,
takes the form

The underlined terms cancel out since (dbt = (bd)- in accordance with
the interchange rule. Here we arrive at result (3). Thus we see, as was
mentioned in Sec. 1.7, that the question of interchanging d and b plays no
principal part for the derivation of the equations. This interchange sim-
plifies the algebra but brings nothing to the final result. This discussion
about the interchange rule could be avoided if the objective was to derive
the differential equations of motion of a system subject to ideal constraints
using the releasing principle.
Equality (3) is as general as the fundamental equations of dynamics. It
presents the result of a formal transformation of the latter and, for this
reason, it is applicable for both holonomic and non-holonomic systems. In
the case of holonomic constraints and independent generalised coordinates
the variations bqs are independent and thus the coefficients in front of
each bqs in eq. (3) must be zero independently. We obtain the system
of differential equations of motion expressed in terms of the generalised
coordinates as
!i 8T _ aT -_ Q 8 (s=l, ... ,n). (7.1.4)
dt 8Qs 8q8
These are Lagrange's equations of the second kind. Their number is equal
to the number of the generalised coordinates, i.e. the number of degrees of
freedom of the holonomic system.
Provided that there are non-holonomic constraints described by the re-
lationships
n
L ak8Qs + ak = 0 (k = 1, ... ,l), (7.1.5)
s=l
7.1 Derivation of Lagrange's equations of the second kind 305

then, by virtue of the theorem of Sec. 6.1, the following equations result as
a consequence of equality (3)

(7.1.6)

These are Lagrange's equations of the second kind in the presence of non-
holonomic constraints. The total number of equations (5) and (6) is n + l
and exceeds the number of degrees of freedom n - l which is the differ-
ence between the number of independent parameters ql, ... , qn describing
the system configuration and the number of equations of non-holonomic
constraints. Equations (5) and (6) have n + l unknown variables consisting
of n generalised coordinates ql, ... , qn and l constraint multipliers AI, ... , A/.
The form (3) of the equation can also be retained with no modification
in the case of redundant coordinates, if the time-derivatives of the finite
relationships between the generalised coordinates are included into eq. (5).
Another derivation of Lagrange's equations is based upon the rearrange-
ment of the left hand sides of equalities (6.3.7). Recalling relationship
(1.3.5) we have

However, due to eq. (1.3.11), we have

d ori OVi
dt oqs oqs'

Thus

and

Taking into account the definition of the kinetic energy (4.1.1) we obtain
N
'""' ori d aT aT
6 mw · · - = - - - - (7.1. 7)
.,=1 "oqs dt iNs oqs '

which is required.
306 7. Lagrange's differential equations

Considering a function of the generalised coordinates, generalised veloc-


ities and time

we introduce the notation

£8 (I) =!!:.. ~f - ~f
dt vqs Vq8
(8 = 1, ... ,n), (7.1.8)

which is referred to as Euler's operator over f. In particular, if f does not


depend on the generalised velocities, then

(7.1.9)

It is clear that

(7.1.10)

Using Euler's operators we cast Lagrange's equations (4) in the form

£8(T)=Qs (8=1, ... ,n), (7.1.11)

and in the case of potential forces accounting for eqs. (9), (10) and (6.4.8)

£8 (L) = 0 (8 = 1, ... ,n). (7.1.12)

where L = T - II denotes the kinetic potential.


When the non-potential forces are present (along with the potential
forces) and some of them are described by the dissipation function <P, La-
grange's equations, by virtue of eqs. (11), (12) and (5.11.12), take the form

(7.1.13)

In the case of non-holonomic constraints (5) Lagrange's equations is writ-


ten as follows

(7.1.14)

The terms

(7.1.15)

in eq. (14) present the generalised constraint forces of the non-holonomic


constraints.
7.2 The energy integral 307

7.2 The energy integral


Lagrange's equations of motion of the second kind is a system of ordinary
differential equations of second order and contains the generalised coordi-
nates, their first and second derivatives with respect to time (termed the
generalised velocities and the generalised accelerations) and possibly the
time t explicitly. This system is linear with respect to the generalised accel-
erations and can be determined as functions of the generalised coordinates,
generalised velocities and time such that
(7.2.1)
An explicit form of these relationships is derived below, cf. eq. (7.4.7).
Let us consider a function of the generalised coordinates, generalised
velocities and time
(7.2.2)
Its time-derivative constructed by means of the equations of motion is given
by

in which the generalised accelerations are replaced by quantities (1) from


the equations of motion, i.e.

d'lj; 8'lj; n (8'lj;. 8'lj;)


-dt = ~ + ~ -;;-qs + ~ fs . (7.2.3)
UL s=l uqs uqs
This function 'lj; of the generalised coordinates, velocities and time is the
first integral of the equations of motion provided that its time-derivative,
constructed with the help of these equations is identically equal to zero.
Then this function 'lj; has the same value for any motion of the system.
Let us construct the derivative of the kinetic potential L with respect to
time by using the differential equations (1.14). We have

The derivatives of L with respect to the generalised coordinates are replaced


here by the expressions using the equations of motion. Now, recalling eq.
(1.5) we can write
dL 8L d n 8L n I n 8<I>
dt = at + dt ~ ~iJ8 -
8=1 q
LQ8iJ8
8=1
+ ~>"kak + ~ 7YiJs
k=l 8=1 qs
308 7. Lagrange's differential equations

or, by virtue of eqs. (6.4.19) and (5.10.3),

(7.2.4)

The equations of motion admit a first integral called the energy integral if all
the active forces are potential, time t does not appear in the expression for
the kinetic potential and the equations for the non-holonomic constraints
do not contain the constant terms ak. Then the right hand side of equality
(4) is zero and we arrive at the following equality

8L
I: a-lis -
n n
I:Pslis - L = L = h, (7.2.5)
8=1 s=l qs

where h is a constant value. The assumption that L does not contain t


explicitly and then all ak = 0 is not equivalent to the assumption of sta-
tionary constraints, since this is also possible in the case of non-stationary
constraints. Below we show examples when L does not contain t explicitly
under nonstationary constraints.
Now we turn our attention to relationship (4.1.13) and recall the expres-
sions for Ps and L. Instead of eq. (5) we have

2T2 + T1 - (T2 + T1 + To) + II = h or T2 + II - To = h. (7.2.6)

This is the expression for the energy integral in the general case. If the
constraints are stationary, then T2 = T, To = 0 and relationship (6) takes
the form

E = T+II = h, (7.2.7)

expressing the law of conservation of the total energy E which is the sum
of the kinetic and potential energies under motion of the system subject to
stationary constraints and only the potential active forces.
Returning to eq. (4) we assume that the constraints are stationary and
that L does not depend explicitly on t (time t may appear explicitly in L
via the potential energy II). Rearranging as above the expression on the
left hand side, we arrive at the relationship

dE =N N* (7.2.8)
dt +,
where N* denotes the power of the non-potential forces except for those
which are expressed by the dissipation function and

N* =- t 8~ lis = -
8=1 8qs
(m + 1) q, (7.2.9)
7.3 The structure of Lagrange's equations 309

denotes the power needed to overcome the dissipative forces. The latter
equality in (9) appears when the dissipative forces are expressed by a single
term of degree m.
The physical meaning of equality (8) is clear: a change in the total me-
chanical energy of the system subject to stationary constraints per unit
time is equal to the power of the non-potential applied forces.

7.3 The structure of Lagrange's equations


Restricting our consideration to the case of holonomic constraints, we write
down Lagrange's equations in the form

£8(L)=Q8 (s=l, ... ,n). (7.3.1)

An extended expression for the kinetic potential can be expressed as follows

L = T - IT = T2 + Tl - (IT - To) = T2 + Tl - IT* , (7.3.2)

where II* depends on the generalised coordinates and, in general, on time

II* = II - To. (7.3.3)

We will refer to II* as the corrected potential energy. As we will see further
on introducing II* has a useful physical interpretation.
Equations of motion (1) can now be written in the form

(7.3.4)

Here T2 means the positive definite quadratic form of the generalised veloc-
ities ql, ... , qn and can be treated as the kinetic energy of a certain system.
For this reason, all terms on the right hand sides of the equations of motion
(4), -£8 (T1 ) included, can be understood as certain generalised forces. Due
to eqs. (1.8) and (4.1.6) we have

The quantities

(7.3.6)
310 7. Lagrange's differential equations

forming the skew-symmetric matrix


o 1'12
1'21 0
1'= (7.3.7)

I'n1 I'n2 0
are named the gyroscopic coefficients whereas
n
fs = L I'skqk (7.3.8)
k=l
denote the generalised gyroscopic force and I'skqk as the gyroscopic forces.
The equations of motion (4) are now cast in the form
all * aBs
£s(T2 )=fs +Qs--a - - a (s=l, ... ,n), (7.3.9)
qs t
in which the latter term vanishes if t does not appear in L explicitly. Only
squares of the generalised velocities appear on the left hand side of eq. (9)
(see Sec. 7.4 for details) and terms linear in the generalised velocities arise
due to the gyroscopic forces, except for those in the generalised forces Qs. In
this case the generalised velocity qs does not appear in the s - th equation,
as I'ss = 0, and the coefficients of qk are equal in magnitude but opposite
in sign to the coefficients of qs in the k - th equation. This property is often
used to prove whether the equations of motion are correctly obtained.
The virtual power (Sec. 5.10) of the gyroscopic forces is equal to zero.
Indeed, by eqs. (8) and (6)
n n n
Nt = Lfsqs = LLl'skqkqs = O.
s=l 8=1 k=l

This explains why the term T1 does not appear in expression (2.6) for the
energy integral.

7.4 Explicit form of Lagrange's equations


Now we write down the right hand sides of the equations of motion (3.4).
Due to eq. (4.1.5) we have
7.4 Explicit form of Lagrange's equations 311

where the latter term, which is linear in the generalised velocities vanishes,
when T does not depend on t explicitly.
Simple transformation yields

where

[k,m;s) = ~ ( 8A kS + 8Ams _ 8A km )
2 8qm 8qk 8qs
denotes Christoffels's symbols of the first kind (introduced in Sec. 4.10) for
matrix A of the coefficients of the quadratic form T.
The equations of motion (3.9) can be recast in the form
n n n
l: Aksqk + l: l: [k, m; s) tiktim
k=1 k=1m=1

rs + Qs - 8II* _ 8Bs _ ~
L..J 8A ks q·k (s = 1, ... , n ) . ( 7.4.1 )
8qs
8 qs k=1 at
As mentioned in Sec. 7.2 these equations present a linear form in the gen-
eralised accelerations and can be resolved for the generalised accelerations
since matrix (4.1.9) is not singular. The inverse A-1 of matrix A was in-
troduced in Sec. 4.2. Their elements are (see (A.3.13))

Asl
-1
= iJ.IAIls (
l, s = 1, ... , n ) , (7.4.2)

where iJ. ls denotes the algebraic adjunct of the element Als of the determi-
nant IAI of matrix A.
Multiplying each equation in eq. (1) by A:;/ and summing up these
products over s we obtain
n n n
ql + l: A:;z1 l: l: [k, m; s) tiktim
s=1 m=1 k=1

(7.4.3)

since
312 7. Lagrange's differential equations

Introducing the values

(7.4.4)

referred to as Christoffel's symbols of second kind (also named the braces)


for matrix A, we arrive at the equations of motion in the generalised coor-
dinates resolved in the generalised accelerations

(7.4.5)

(l=I, ... ,n).

In the case of stationary constraints Lagrange's equations (1) and (5)


simplify and take the form

all
L +L L
n n n
A k8ijk [k, m; s] qkCim = Q8 - [ ) (s = 1, ... ,n), (7.4.6)
k=l k=lm=l ~

ij8 + ~tl {k~}qkqm = ~A~l (Ql - ~~) (s = 1, ... ,n).


(7.4.7)

7.5 Geometric interpretation of particle motion


Lagrange's equations in the form (4.6) and (4.7) admit an interesting and
fruitful interpretation in terms of tensor analysis and Riemann's geometry.
The formalism needed for understanding this can be found in Appendix
B. We begin with the simplest case which is motion of a free particle.
Instead of the Cartesian coordinates x and y it is convenient to introduce
the variables

x' = Vmx, y' = Vmy, Z' = Vmz, (7.5.1)

where m denotes the mass of the particle. Let us enter the new position
vector, its velocity and acceleration which differ from the true ones by the
proportionality factor Vm, such that

(7.5.2)
7.5 Geometric interpretation of particle motion 313

Here is denote the unit vectors of the Cartesian coordinate system Ox' y' z'.
The expression for the kinetic energy is as follows

T = ~ v2 = ~
2 2
(dS)
dt
2
'
(7.5.3)

where ds is the" arc element" of the particle trajectory. Introducing the


curvilinear coordinates ql, q2, q3

(7.5.4)

we obtain
• • s
r=v=rsq, (7.5.5)
r s denoting the basis vectors, see (B.4.4). The redundant summation sign
is omitted due to the convention of summation over the repeated index s.
The generalised velocities thus are contravariant components of the velocity
vector v. The expression for the kinetic energy in terms of these components
is given by

(7.5.6)

The coefficients of the kinetic energy (earlier denoted as Asd play the role
of the covariant components of the metric tensor. This enables us to write

(7.5.7)
The elementary work of a force F applied to the particle due to a virtual
displacement from the position under consideration is

, 8r 1 8 s Q8 s
8W=F·rm=rmF.rsq = sq· (7.5.8)

We introduce the force vector

(7.5.9)

According to the above definitions Qs are the generalised force and the
covariant components of force Q. The contravariant components are given
by
(7.5.10)

We proceed now to construct an expression for w. Differentiating eq. (5)


with respect to time and making use of rule (B.4.9) for differentiation of
the basis vectors, we obtain
314 7. Lagrange's differential equations

The quantities in parentheses are the contravariant components WS of


vector w

w s = q··s + { tks } qq.


·t ·k (7.5.12)

Under the introduced notation for vectors wand Q Newton's second law
takes the form

w=Q, (7.5.13)

that is the acceleration is equal to the force applied to the particle. By meanS
of eqs. (10) and (12) we have

q··s + { tks } q·t q·k -_ QS -_ 9 skQ k· (7.5.14)

Hence we have Lagrange's equations in form (4.7) resolved in terms of the


generalised coordinates. Lagrange's equations in form (4.6) and its COnven-
tional form (1.11) are obtained by using relationships (13) in terms of the
covariant components.
Indeed,

Ws = gskWk = gsk ( q..k + {k}


mt q·m q.t) = gskq··k + [m, t; sJ qq.
·m ·t (7.5.15)

Taking into account that

··k d .m ogsm.m ·t
gskq = dt gsmq - oqt q q, ( ogsm
oqt
_ [ t.
m" s
J) q·m·tq -_ .!.2 ogmt ·m ·t
oqS q q

and using expression (6) for the kinetic energy we obtain


d oT oT
W ----- (7.5.16)
s - dt oi]s oqs .
Thus, Newton's second law in the covariant components takes the form
of Lagrange's equations

~ oT _ oT = Qs (7.5.17)
dt oi]s oqs
or, in the explicit form (4.6),

gskq··k q q = Qs·
+ [m, t ; s J.m.t (7.5.18)

Now we proceed to the equations of motion in the natural form. The


velocity vector (5) can be written in the form

.k dqk.
V = rkq = rk ds s, (7.5.19)
7.5 Geometric interpretation of particle motion 315

where ds is the arc elements defined by eq. (7). The vector

(7.5.20)

having the contravariant components dqk / ds is directed along the trajec-


tory tangent. It follows from eq. (7) that T is the unit vector. Then we
obtain

• .. .2 dT
V=TS, W=V=TS+S ds. (7.5.21)

Now we have

dT = rk d 2 qk + dqk ark dqm = rk (~qk + { k }dqm d qt ) = kn.


ds ds 2 ds aqm ds ds 2 mt ds ds
(7.5.22)

This vector is referred to as the curvature vector. The quantities

are its contravariant components whereas the quantity

(7.5.23)

is the curvature of the trajectory. The unit vector n represents its main
normal. Hence, we obtain

(7.5.24)

The acceleration vector is represented as a sum of two components,


namely the tangential W(T) and normal W(n) accelerations

(7.5.25)

Decomposing vector Q in a similar way

(7.5.26)

we obtain by eq. (20)

(7.5.27)
316 7. Lagrange's differential equations

and furthermore

The natural equations of motion are thus written in the form

.. _ Q dqt 2qt +{ t }dqm d qk ) 82 = Qk (gkt _ dqt d qk ) .


s- tds' (d 2
ds mk ds ds ds ds
(7.5.29)

In the case of the potential forces the first term can be transformed as
follows
.. d8 .
s= - s
d 82
= -- = - - - =--
oil dqt dil
ds ds 2 oqt ds ds .

Integration yields

82 = 2 (h - II) , (7.5.30)

where the integration constant h is the total energy of the particle. Elim-
inating 82 from eq. (30) and the second equality in eq. (29) we arrive at
the following differential equations for the trajectories of a particle in the
potential field

d 2 qt +{ t }dqm dqk = (7.5.31)


ds 2 mk ds ds

Of course, they must be considered along with the following relationships

(7.5.32)

7.6 Motion of a particle on a surface


The simplest example of motion of a system subject to constraints is the
motion of a particle on a surface. The equation of the surface is assumed
to be given in the form (cf. (B.7.1))

(7.6.1)

where qCi denote the Gaussian coordinates. In what follows, the Greek
indices take values of 1 and 2. The velocity

v = PCiljn (7.6.2)
7.6 Motion of a particle on a surface 317

lies on the surface, however the time-derivative of the velocity vector does
not. Applying formulae (B.7.19) we find

(7.6.3)

where m denotes the unit vector of the normal to the surface and bo. {3 stands
for the coefficient of the second quadratic form of the surface. Christoffel's
symbols are calculated by means of the metric tensor whose covariant com-
ponents ao.{3 are the coefficients of the first quadratic form of the surface

(7.6.4)

with T being the kinetic energy of the particle. The vector v implies the
total acceleration.
Formula (3) presents the total acceleration as a sum of two vectors. The
first one is the vector of acceleration on the surface w with the contravariant
components

(7.6.5)

obtained by covariant differentiation of the contravariant components of


the velocity. The second one is the acceleration normal to the surface

_ b .0. .(3 _·2 b dqo. dq{3 _ .2k-


W(m) - m o.{3q q - s m o.{3 ds ds - s m. (7.6.6)

Here, due to (B.8.4), the value of k denotes the normal curvature


k= bo.{3dq o. dq{3
(7.6.7)
ao.{3dq o. dq{3·

The covariant components of the accelerations on the surface can be


rearranged into the form

(7.6.8)

Introducing the independent variable s in expressions (5) for the contravari-


ant components w o. of the acceleration vector w on the surface we obtain

(7.6.9)
318 7. Lagrange's differential equations

where k*a denote the contravariant components of the vector of geodesic


curvature defined by eqs. (B.8.6). Its accompanying unit vector n* deter-
mines the direction of the geodesic normal of the trajectory. Now we have

(7.6.10)

where the geodesic curvature k* is given by eq. (B.8.8). The vector of ac-
celeration on the surface is the sum of two vectors, namely the tangent ac-
celeration and the geodesic normal accelerations. According to eq. (B.8.13)
they are mutually perpendicular.
Let us consider the active force F and the reaction force of the surface
R, the latter being perpendicular to the surface in the case of no friction,
acting on the particle. The elementary work of these forces is

(7.6.11)

As before, we introduce the vector Q with the covariant components Qa


which are equal to the generalised forces. However equality (5.9) is no longer
valid since the active force F may have a component along the normal to
the surface

(7.6.12)

Newton's second law now takes the form

v= Jrn (F + R) .
First, we arrive at the equations of motion

(7.6.13)

or, in the covariant components,

d aT aT
dt aqa - aqa = Qa. (7.6.14)

Second, we obtain the expression for the normal reaction force

(7.6.15)

the presence of the factor Vm being explained by the length scaling due to
eq. (5.1).
The equations of motion of the particle on the surface coincide with those
for a free particle. If one "forgets" the reaction force of the surface and the
component of the active force normal to the surface, the equations (13) and
7.7 Examples 319

(14) can be treated as the second law in the form of (5.13) for the vectors
of the force and acceleration on the surface.
Repeating the derivation of the previous section we can write the equa-
tions of motion in the form corresponding to the natuml equations of mo-
tion in projections on the tangent and geodesic normal of the trajectory

2qOt
(d 2
ds
+ {a }dqf3
i3"f ds ds
:q~)~: ~~ (af3Ot _ qf3 qOt ). }
{3
d d
ds ds
(7.6.16)

If Q = 0, the velocity remains constant. We will refer to such a motion as


inertial. The inertial motion, as it follows from eq. (16), takes place along
the geodesic line of the surface.

7.7 Examples
7.7.1 Motion of a free particle relative to a non-orthogonal
coordinate system
Let us construct the equations of motion of a heavy particle relative to
a system of rotating axes whose origin moves vertically with a constant
velocity, cf. example in Sec. B.6.
Using formulae (5.12) and expressions (B.6.6) for Christoffel's symbols
we obtain the contravariant components of the acceleration in the form

X
··1
+ 2 { 231 } x·2 x·3 + { 331 } x·3 x·3 = x··1 - 2 . 2 ·3
TX X -
·1 2 (.3) 2
X T X ,

5;2 + 2Tj;1j;3 _ j;2T2 (j;3) 2 ,

5;3,

where T denotes a constant coefficient relating the angle of rotation of the


axes to the vertical displacement.
Provided that the motion takes place in the field of gravity and axis Oz
is directed vertically then for a particle of unit mass we have

n = gz = gx3,
and the covariant components of force Q are as follows

Q3 =- aan
x3 = -g.
The contravariant components are equal to
320 7. Lagrange's differential equations

The equations of motion take the form

Xl - 2Tj;2j;3 _ XlT2 (j;3)2 = -gTX 2 }

X2 + 2Tj;Ij;3 _ X2T2 (j;3)2 = gTXI, (7.7.1)


X3 = -g.

In order to write down the equations of motion in terms of the covariant


components we construct the expression for the kinetic energy

Then we obtain

(7.7.2)

It is easy to prove that the projections of acceleration w of the particle


On the rotating axes Oxl and OX2 are equal to the covariant components
WI and W2, whereas the projection on axis Oz is equal to the contravari-
ant component W3' The complexity of the obtained system of equations is
explained by the choice of the coordinate system. Taking a fixed coordi-
nate system, the motion describes the fall of a heavy particle in a vacuum.
Provided that the trajectory plane (the plane in which the vector Vo of
the initial velocity lies) coincides with plane 0 zx and the origin of the
coordinate system is placed in the initial position of the particle, then we
have

1 2
Z = vozt - -gt
2

and furthermore

3 1 2
X = vozt - -gt
2

Clearly, these equalities satisfy eqs. (1) and (2) identically.


7.7 Examples 321

7.7.2 Equations of motion of a free particle relative to an


orthogonal coordinate system
The equation of motion of a free particle (5.18) for the coordinate ql refer-
ring to a orthogonal curvilinear coordinate system is given by

(7.7.3)

where hi denote Lame's coefficients. The other two equations are obtained
by an appropriate change of the indices. Expressions (B.1O.8) for Christof-
fel's symbols are used here. Equations in terms of the contravariant com-
ponents, i.e. in the form of eq. (5.14) are obtained by dividing eq. (3) by
coefficient hi.
As an example let us consider an orthogonal system of curvilinear coor-
dinates ql = a and q2 = (3 lying in the plane Oxy and forming a network
of isometrics. This means that the Cartesian coordinates can be expressed
in terms of a and (3 as follows

x + iy = f b) = f (a + i(3), (7.7.4)

where f b) is a function ofthe complex variable 'Y = a + i(3. Its derivative


f' b) is assumed to be non-zero in the domain of the complex variable 'Y.
Then

dx + idy = f' b) (da + id(3) , ds~ = If' b)1 2 (da 2 + d(32) .

Therefore, Lame's coefficients for the isometric coordinates ha and h{3 co-
incide and are equal to

h~ = h~ = If' b)1 2 = f' b) l' COy) = h 2 (a, (3), (7.7.5)

the bar denoting the complex conjugate. A spatial coordinate system is


obtained by rotating plane Oxy about axis Ox. The square of the arc
element is then given by

where cp denotes the azimuthal angle and

(7.7.6)
322 7. Lagrange's differential equations

The equations of motion (3) in terms of the coordinates a, j3, VJ have the
form

(7.7.7)

where it is assumed that Q<p = O.


The latter equation yields the integral of area

(7.7.8)
where k<p is an integration constant. Removing cp by means of this relation-
ship and assuming the existence of potential forces we obtain

a
h(. 2
.. + a;;-
aIn a - j3
8ln h..
.2) + 27ij3aj3 = -
1 a
h2 aa
k~))
( II + 2h~ ,
(7.7.9)
.. alnh
j3 + aj3
(.2j3 _a.2) + 2 8lnh . . __ ~~ (
oa aj3 -
k~ )
h2 o{3 II + 2h~ .

Now, using the energy integral

h2 ( a 2 + l) = 2(
Eo - II - 2~~) , (7.7.10)

it is easy to rearrange the first equation in (9) into the form

.. . (lnh2)- = 1- -
a+a
h 2 oa
[0 ( Eo-II--
2h2<p
k~) +aIn h
- -2 ( E o - I I -k~-
oa 2h2<p
)1
.

The second equation can be written by analogy.


Hence we obtain the system of equations

(7.7.11)

admitting further integration if the quantity in the parentheses on the right


hand side is a sum of two functions, each depending on a and j3 separately.
This is, for instance, the case for planar motion of a particle attracted by
7.7 Examples 323

two centres. Placing these centres Fl and F2 on axis Ox and denoting the
distance between them by 2c, we will determine the position of the moving
particle M using the curvilinear coordinates introduced by the following
relationships

x + iy = c cosh (a + i,8) = c (cosh a cos,8 + i sinh a sin,8) .


Then

The distances Tl and T2 from particle M to the attracting centres are found
using the equalities

Tl c Icosh (a + i,8) + 11 = c (cosh a + cos,8) ,


T2 c Icosh (a + i,8) -11 = c(cosha - cos,8).

Taking into account that the potential energy is given by

IT = _/1 - 12 = - 1 [(/1 + h) cosh a - (11 - h) cos,8J,


Tl T2 cosh 2 a - cos 2 ,8

where /1 and 12 are constant values and kcp = 0 for the plane motion we
obtain

h 2 (Eo - IT) c2 ( Eocosh 2 a+ /1+12


c )
cosha-

c 2(Eo cos2,8 + /1-12


c cos ,8 ) .

Integration of equations (11) yields

h 4 ix 2 2c2 (Eo cosh 2 a + /1 : 12 cosh a + 'h) ,


h4~2 -2c2 (Eo cos 2 ,8 + /1 : 12 cos,8 + "12) .

Adding these equalities and using the energy integral is it easy to find that
"11 + "12 = O. This we arrive at the system of two equations

2
Eo cosh a +
/1+12 cosh a - "12 -Eocos2a
fJ-
II-h cos,8+"12
c c
2 (dt)2
c2 (cosh2 a - cos 2 ,8 )2'
324 7. Lagrange's differential equations

which yields another two integrals

J( Eo cosh 2 a +
1 +1
1 C 2 cosh a -1'2
)-1/2
da-

J( -Eo cos 2 f3 -
1 f
1: 2 cosf3 + 1'2
)-1/2
df3 = C,

t+ to = v'2 [
C
J cosh a
2 ( Eo cosh 2 a +
11+ f 2
c cosh a -1'2 )
-1/2
da-

J f3 (
cos 2 -Eocos2 f3- 11: 12cos f3 +1'2 )-1/2 df3. 1
(7.7.12)

The solution of the problem having four integration constants to, C, Eo, 1'2
is thus reduced to calculating the above elliptic integrals. In passing we no-
tice that the detailed investigation of motion of a particle in the field of
two attracting centres was given in the treatise on celestial mechanics by
Charlier, see [19]. The problem of two attracting centres belongs to those
problems whose reduction to quadratures was pointed out by Liouville. The
systems for which

(7.7.13)

are named the systems of Liouville's type. A more complete definition of


Liouville's systems will be given in Sec. 10.14. Returning to eq. (11) and
repeating the calculation, we obtain for Liouville's systems

J( Eoh12 - WI - l' )-1/2 da - J( +


Eoh22 - W2 l' )-1/2 df3 = C,
v'2 (t + to) = Jhi (Eohi - WI -1') -1/2 da+

Jh~ (Eoh~ - W2 +1')-1/2 df3.


(7.7.14)

7.7.3 Equations of motion on the surface of revolution


These equations have the form
.. dr .2 dr 1
S = -r ds 'P = Qs, 0+2-<jJr
r ds
= -Q
r <P
. (7.7.15)

We use the notation of Sec. B.10 as well as Christoffel's symbols calculated


there. If the axis of the surface of revolution coincides with the vertical axis
and the motion takes place in the gravitational field then
dz
Qs = -g ds' Q<p = O.
7.7 Examples 325

Integrating the second equation and eliminating 02 from the first equation
we obtain
.. 1 4.2 d 1 dz
2· 2·
r ifJ = roifJo, s + -roifJo--
2 ds r2
= -g-.
ds
(7.7.16)

The latter equation yields the energy integral

(
82 = 2 h - gz _ roifJo
4.2) . (7.7.17)
2r2
Here z and r are viewed as functions of the arc measured along the meridian
and h is the integration constant. The problem is thus reduced to quadra-
tures. In the case of a spherical pendulum

z = R cos iJ, r = R sin iJ = viR2 - Z2,

and the obtained differential equation


z2 = 2 (h - gz) (R2 - z2) - r;506
can be integrated in terms of the elliptic functions. The motion of a particle
on the surface of a circular cone

z = scosa, r = ssina,
the surface of a paraboloid of revolution

and other surfaces can be reduced to elliptic functions, see [95].

7.7.4 Motion on a developable surface


We consider now the case of motion of a particle on a developable sur-
face. Since the arc length and the geodesic curvature of the trajectory are
the invariants of the surface, they remain unaltered when the surface is
developed on the plane. For this reason, the equations of motion on the
developable surface are written in the form of equations of motion of the
particle subject to an active force that is equal to the force component in
the plane tangent to the surface.
Let us consider, for instance, a heavy particle moving on the surface of
a right circular cone with a vertical axis. The particle is assumed to be
subject to the force 9 cos a directed along the generator of the cone. When
the cone is developed on the plane we arrive at the problem of planar motion
of the particle subject to a central force of constant value, the centre of the
force coinciding with the cone vertex. The equations of motion in polar
coordinates are as follows
·2 d 2'
S - s'lj; = ~gcosa, dtS 'lj; = 0,
326 7. Lagrange's differential equations

where the negative sign (force of attraction) corresponds to the lower po-
sition of the cone vertex and the positive sign (force of repulsion) to the
upper position of the vertex. The angles 'lj; and 'P (the azimuth on the cone
surface) are related by 'lj; = 'Psina. The result is
4. 2
82 = 2 (h =fgscosa) _ so~o.
s

7.8 Geometrical interpretation of the equations of


motion of the system
Our consideration is restricted to the case of a system of particles subject
to stationary holonomic constraints. A review of papers on the geometric
methods in dynamics can be found for example in [86]. Similar to Sec. 7.5 we
determine the position of particle Mi of mass mi by quantities proportional
to the Cartesian coordinates of the fixed basis system

~3i-2 = VmiXi, ~3i-l = VmiYi, ~3i = VmiZi, (i = 1, ... ,N).


(7.8.1)
We can say that the position of the system at a given time instant is
described by a point (~1' ... '~3N) of the Euclidean space E 3 N. The position
vector of this representative point is denoted by r, and quantities ~v are
understood as its projections on axes of the Cartesian coordinate system
in E 3N . The relationship
(7.8.2)
expressing the Cartesian coordinates in terms of the generalised coordinates
q\ ... , qn defines a Riemannian manifold Rn in E 3N . The problem is to
describe the motion of the material system by means of Rn. We will repeat,
in a certain sense, the content of Sec. 7.6 where the motion on a surface is
studied. However one should not forget that the words "position vector",
"velocity" and "acceleration" imply concepts related to the representative
point and not to a specified particle of the system.
The coordinate basis at any point of Rn is determined by the system of
n vectors
8r
r", = 8q"" (7.8.3)

where the Greek indices take values 1, ... , n. The vector represented as a"'r",
belongs to Rn and a'" denote the contravariant components of this vector.
Now, differentiating eq. (2) with respect to time we obtain the velocity
vector
(7.8.4)
7.8 Geometrical interpretation of the equations of motion of the system 327

belonging to Rn. Here il' denote its contravariant components.


The kinetic energy of the system of particles is given by

T
N

= "21L--
" (.1 ·2 .2) 1 1.. 1 2 "c
mi Xi + Yi + Zi ="2 L-- 'w = "2 r . r ="2 V .
3N
2

i=1 v=1

Expressing v by means of formula (4) we obtain

T 1 .f3 1 .f3
= "2 roo . rM = "2 aooM
·00 ·00
q q , (7.8.5)

where a oo f3 denote the contravariant components of the metric tensor of


space Rn.
The active forces F i acting on the particles are replaced on aggregate
by the vector Q termed the force related to the representative point. The
vector belonging to Rn is determined by its covariant components which
are equal to the generalised forces

Q = Qoor OO , (7.8.6)

rOO being the vectors of the co-basis. The above concept of the force is due
to the fact that the elementary work of all active forces Fi due to virtual
displacement of the system particles is equal to the elementary work of
"force" Q due to virtual displacement 8r

Q. 8r = QoorOO . rf38qf3 = Qooa38qf3 = Qoo8qOO.


The acceleration vector is defined as the following vector

(7.8.7)

belonging to Rn and having the contravariant components WOO. They are


obtained by covariant differentiation of the contravariant components qn
of the velocity vector v with respect to time

w
00 _ ··00
- q
+ { (3"(
a } qq.
.f3 .'"'( (7.8.8)

Referring to the case of motion of a particle on a surface it can be said that


w represents that part of vector v which belongs to Rn. It is worth paying
attention to the concluding remark in Sec. B.12.
The contravariant components WOO are calculated due to the known rule

(7.8.9)

Repeating the calculation of Sec. 7.5 we can write this expression in the
form
00 doT oT
w =------ (7.8.10)
dt oqOO oqOO·
328 7. Lagrange's differential equations

Based upon these definitions we can now formulate the law of motion
in the form of Newton's second law, namely that the acceleration vector is
equal to the force vector

w=Q. (7.8.11)

Using the contravariant components we obtain

•• <>
q
+ { i3'Y
a } .{3 ."1 _
q q - a
<>{3Q (3. (7.8.12)

The result is the equations of motion in the form (4.7) resolved in the gen-
eralised accelerations. Using the covariant notation we obtain Lagrange's
equations of the second kind. Thus, the latter express Newton's law for the
motion of the particle representing the system of particles under considera-
tion in space Rn with the metric determined by the quadratic form 2Tdt 2 .
The laws of motion thus become a geometrical interpretation. Repeating
the derivations of Secs. 7.5 and 7.6 we can write the equations of motion
in the form of the natural equations which is immediat€ly derivable from
eq. (5.29)

.. _Q dq<>
s- <> ds'
82 (d 2q <>
ds 2
+ {ai3'Y }dq{3 dq'Y)
ds ds
= (3) Q
(a<>{3 _ dq<>dq
ds ds {3
.

(7.8.13)

In the case of motion by inertia, i.e. Q = 0, the motion occurs with a


constant velocity along the geodesic lines of the manifold Rn for which

d2 q<> +{ a } dq{3 dq'Y = o. (7.8.14)


ds 2 i3'Y ds ds
The sequel of the natural equations is Synge's generalisation of Bonnet's
theorem on the motion of a particle, see [95). Let the representative point
have the same trajectory C under any of the forces QCl)' Q(2)' ... , QCm)
acting separately. The kinetic energies of each of these separate motions are
denoted as TCl), T(2)' ... , TCm). Then the same trajectory is realised when the
above forces Q(1), Q(2)' ... , QCm) act simultaneously and the kinetic energy
of this motion is simply the sum of the kinetic energies of the separate
motions.
Indeed, accounting for eqs. (5.24) and (5.22) we can write the natural
equations of motion in the form

dTCi)
~T + 2TCi)kn = QCi) (i = 1, ... ,m), (7.8.15)

where T and kn are the same for any i. Addition of these relationships
yields the required result.
7.9 Example of applications of Lagrange's equations 329

7.9 Example of applications of Lagrange's


equations
7.9.1 Double mathematical pendulum in the case of a moving
suspension point and a quadratic resisting force
Figure 5.14 shows the system. The motion of two bodies in the vertical
plane is considered. The bodies are assumed to be so small that they can
be viewed as particles Ml and M 2. The lengths of the inextensible strings
OM1 and MIM2 are hand b, respectively. The suspension point 0 moves
in the vertical plane OM1M 2. The velocity of point 0 is denoted by Vo
and the angle between Vo and the downward vertical is denoted by 0;. The
resisting force of the motionless air is assumed to be proportional to the
square of the particle velocities. We require the equations of motion taking
into account force due to gravity.
The calculation of the dissipation function and the generalised resisting
forces was performed above in Sec. 5.12. Equations (5.12.2) provide us with
expressions for the squares of the particle velocities. This enables the kinetic
energy to be expressed as
1
T = "2 [(ml + m2) l?tPi + m2l~tP§ + 2m2hl2tPltP2 cos (<PI - <P2)] +
1
Vo [(ml + m2) htPl sin (0; - <PI) + m2l2tP2 sin (0; - <P2)] + "2 (ml + m2) v6
= T2 +Tl +To·
(7.9.1)

The components Tl and T2 are due to the non-stationarity of the con-


straints. We calculate Euler's operators

El (T2) = (ml + m2) li({Jl +m2h l2({J2 cos (<PI - <P2)+m2h l2tP§ sin (<PI - <P2)
E2 (T2) = m2l§({J2 + m2h l2({Jl cos (<PI - <P2) - m2l1l2tPi sin (<PI - <P2) .

While calculating Es (Tl) the quantities Vo and 0; are considered as pre-


scribed functions of time. The result is

The gyroscopic terms do not appear since the coefficient of tPl depends
only on <PI' and that of tP2 only on <P2. The term To does not depend
on the generalised coordinates and the corresponding Euler operators are
identically equal to zero.
The potential energy of the gravity force is

(7.9.3)
330 7. Lagrange's differential equations

and the equations of motion have the form

(m1 + m2) l~CP1 + m2h l2CP2 cos (CP1 - CP2) + )


m2hl2cp~ sin (CP1 - CP2) = - (m1 + m2) h [vo sin (a - CP1) +
voa cos (a - CP1)] - 9 (m1h + m2l2) sin CP1 + Qb
m2l~cp2 + m2hl2CP1 cos (CP1 - CP2) - m2hl2cp~ sin (CP1 - CP2)
= -m2l2 [vo sin (a - CP2) + voacos (a - CP2)]- gm2l2 sinCP2 + Q2,
(7.9.4)
the expressions for the generalised resisting forces being given by formulae
(5.12.4).
We restrict out attention to the case of uniform horizontal motion of the
suspension point and small deflections of particles M1 and M2 from the
position of relative equilibrium. Then, setting in eq. (4)
7r
Vo = 0, a="2'
and making use of expressions (5.12.6) for the generalised resisting forces
and equations for the relative equilibrium we arrive at the system of linear
differential equations with constant coefficients

aUE1 + a12E2 + bUt1 + b12t2 + C1C1 = 0, } (7.9.5)


a21E1 + a22E2 + b21 t 1 + b22t 2 + C2C2 = 0,
where

au

bu

Here ->'1 and ->'2 denote the tangents of the angles of the strings in the
position of relative equilibrium

Assuming the absolute values of these angles are smaller than 90° we have
o 1 o 1
COSCPl = ~' cosCP2 = ~.
VI + >.~ VI + >.~
7.9 Example of applications of Lagrange's equations 331

Z M

J1LMm>
FIGURE 7.1.

The differential equations (5) can be viewed as Lagrange's equations


constructed by means of the following quadratic forms for the kinetic and
potential energies and the dissipation function

It is easy to prove that these quadratic forms are positive definite. The
equation of energy (2.8), which in this case takes the form
d
dt (T* + II*) = -2<1>*,
indicates that the total mechanical energy of the system decreases mono-
tonically.

7.9.2 Motion of a folded string


We consider the folded inextensible string shown in Fig. 7.1 in which a
particle M of mass m moves along the right part of the string with a
constant relative velocity. This problem was considered by Hamel in [36] in
which it was assumed that the mass m was absent. We assume the initial
conditions which ensure that the folded string moves vertically as the right
part shortens and the left part lengthens. Let x and y denote the height
of the right and the left ends of the string at time instant t, respectively,
and z denote the height of the point K at which the string is folded. The
length of the string is l.
As follows from Fig. 7.1
z- x +z - Y = l,
332 7. Lagrange's differential equations

ensuring that the system has two degrees of freedom.


Introducing the quantity

u=x-y (7.9.6)
one obtains the following
1 1
z=x+ "2 (Z - u) = y + "2 (Z + u). (7.9.7)

As the string is inextensible, the velocity of each point of the right part
is x and that of the left part is y. Thus the kinetic energy of the system is
T ~p[(z-x)x2+(Z-Y)1?] +~m(x+~)2
~ p [2ZX2 - 2ux (Z + u) + if? (Z + u)] + ~ m ( x + ~ f.
Here p denotes mass of length unit and ~ denotes the distance from point
M to the end of the right part at time t.
The potential energy due to gravity is given by

II = pg [~(z+x)(Z-x)+~(Z+Y)(Z-Y)] +mg(x+~)
pg (zx - ~zu - ~u2) + mg (x +~) + const.
Taking into account that ~ = canst we obtain the differential equations

(x + g) (1 + : ) = ;Z [u (I +lu) + u 2 ], }
(7.9.8)
(x + g) (1 + u) = U (1 + u) + "2U2.
Removing x + 9 we arrive at the following equation

..
U ="21 (12m
Z - Z
1 ) u.
+u .2 (7.9.9)
+--u
p
It has a particular solution u = 0 corresponding to the case in which both
end of the string have initial velocities Xo = Yo. It follows from eqs. (6) and
(8) that x = ii = -g. The string moves vertically with a constant acceler-
ation and the length of each part does not change. This trivial solution is
of no interest.
Let us return to eq. (9) and take u as an independent variable, thus

.. du.
u=-u.
du
7.9 Example of applications of Lagrange's equations 333

As it =1= 0, eq. (9) admits a separation of variables

dit 1 du 1 du
-it = - 2-l+-u + 2-l----,2""""m--
+--u
p

Let us assume that the length of each part of the string is l/2 at t = 0,
while the difference between the initial velocities of the right and left parts
is positive, i.e.

t = 0, u = 0, it = ito > 0.

Integration yields

it = ito (7.9.10)

It is clear that the value of u cannot exceed l. Moreover, u increases mono-


tonically from zero and reaches 1 at the time instant when the string
straightens itself.
An expression for t in terms of it is easily obtained by integrating eq.
(10)

J
u

itot= (l + y) (
1- U2m) duo (7.9.11)
o l+-
P

This integral can be easily evaluated, however its expression in closed


form is of no interest.
Now we proceed to calculating the tension in the different parts of the
string. The differential equation of motion for the right part of string having
length si < ~ is as follows

ps~x = S~ - ps~g or ps~ (x + g) = S~,

where S~ is tension in the lower right part.


Considering a section of the right part of length s~ > ~, it is necessary
to take into account mass m. Then

ps~ X + mx = S~' - ps~ 9 - mg or S~' = (ps~ + m) (x + g) .


334 7. Lagrange's differential equations

Replacing x+ 9 by means of (8) and considering relationships (9) and (10),


we obtain

(7.9.12)

For a section of length 82 of the left part of the string

..) (.. ..) 1'2 Z(Z+7)


S
2 = P8 2 (Y + 9 = P8 2 X +9 - u ="2 pU 082 2 ( 2m ) .
(Z+U) Z+--U
p
(7.9.13)

By virtue of (7) one obtains that at point K


1
81" =Z - x ="21 ( Z - U) , 82 = Z - Y= "2 (Z + u)
and

(7.9.14)

The tension at all cross-sections in the string is positive at any time


instant which implies that the motion considered is feasible. 1

7.9.3 Gyroscope in a Cardan suspension


An expression for the kinetic energy of the system consisting of the rotor,
the gimbal and the outer ring was constructed in Sec. 4.12. A point mass m

1 As shown in a recent paper by W. Steiner and H. Thoger On the equations of motion


of a folded inextensible string, Z. Angew. Math. Phys., Vol. 46, 1995, pp. 960-970, the
use of Lagrange equations in their conservative form may result in incorrect equations
of motion. This follows from the energy loss at the moving fold in the case of a zero
radius of the fold. Yu.G. Ispolov derived the equations of motion for a string moving in
a massless U-tube by means of the balance of momentum and angular momentum. The
above equations for the different parts of the string are proved to be correct, even for
the case of a vanishing radius of the fold.
7.9 Example of applications of Lagrange's equations 335

i«I

J I
i'J
(if
'I
(- )
If
I

!lIn
b)

FIGURE 7.2.

is attached to the gimbal (i.e. the inner ring) at the point where the rotor
axis intersects the gimbal. The aim here is to construct the equations of
motion in two cases: a) the axis of the outer ring is horizontal; and b) it is
vertical. The corresponding initial positions are depicted in Fig. 7.2a and
b.
It is necessary to add the kinetic energy of the mass to the derived
expression of the kinetic energy (4.12.1). The velocity vector of this point
is

where h denotes the distance between this point and the point of intersec-
tion of the suspension axes. The kinetic energy of mass m is thus

(7.9.15)

which implies that the only correction is to add mh 2 to the moments of


inertia A2 and B2 of the gimbals. We denote A; = A2 + mh2, B~ =
B2 +mh2 .
The position of the axes is shown in Fig. 7.3. Axes O~'T]( denote axes
attached to the fixed base of the facility, axis O~ being the axis of rotation
of the outer ring. The trihedron of unit vectors i~ , i; , i~ is related to the
inner ring. In the case a) axis O~ is horizontal whereas in case b) it is
vertical. The potential energy of the first case is
II = mg( = mgh cos a. cos (3,
while in the second case
II = mg~ = mgh sin (3.
This follows from the matrix of the directions cosines (2.6.2) and can be
easily proved with the help of Fig. 7.3.
336 7. Lagrange's differential equations

FIGURE 7.3.

Applying eq. (1.12) we obtain Lagrange's differential equations for the


system considered

(al + b cos2 13) Ci - ba/3 sin 213 + C3 (<p + a sin 13) /3 cos 13+
C3 sm. 13 ( . . . 13)- = - an
r.p + 0: sm 00: '
(7.9.16)
a2 13·· + 2 • 213
1 b0:. 2 sm . 13)·0: cos 13 = -
- C3 (.r.p + 0:. sm an
013'
C3 (<p + asinj3r = 0,

where

The latter equation admits the integral

(7.9.17)

expressing the conservation of the kinetic momentum of the rotor which


is the projection of the resultant angular momentum about fixed point O.
Using this integral, we can write eq. (16) in the form

(al + b cos2 13) Ci - baj3. sin 213 + C3roj3. cos 13 = - an


00: ' }
.. 1 2 an (7.9.18)
a2j3 + 2ba sin 213 - C3r oa cos 13 = - 013 .

The terms -c3ro/3 cos 13 and C3roa cos 13 play the part of the generalised
gyroscopic forces. Though the constraints are stationary and term Tl is
absent from the expression for the kinetic energy the above terms appear
because of elimination of the generalised velocity <p from the equations of
motion. This topic is discussed in detail in Sec. 7.17.
7.9 Example of applications of Lagrange's equations 337

The integral of energy (2.17) has the form

T + II = ~ [(a 1 + bcos2 (J) a? + a2il] + II = Eo, (7.9.19)

the term ~C3r~ being included in Eo.


In the case of the vertical axis of rotation, the system admits one more
first integral. In this case II is independent of a and since a does not appear
in the expression for the kinetic energy, the first equation in (18) can be
written as follows

yielding the following integral

~~ = (al +bcos2{J)a+C3rosin{J=Kr:,. (7.9.20)

Here Kr:, denotes projection of the resultant angular momentum KO on


axis Of The projection conserves its constant value as the moment of the
external forces, that is the weight and the reaction forces at the bearings
of the outer ring, is equal to zero about the vertical axis of this ring.
Returning to case a) we assume that the deviation of the ring from the
initial position shown in Fig. 7.2a remains small. Then the angles a and {J
are small and the linearised equations of motion (18) are
(7.9.21)
Following the standard procedure, a particular solution of this system of
first order linear differential equations is sought in the form

a = Dl cos (kt + 'lj;), {J = D2 sin (kt + 'lj;). (7.9.22)


We obtain two linear homogeneous equations in Dl and D2

The nontrivial solution exists when the determinant of the system is zero,
i.e.

(7.9.23)
If h < 0 which means that the mass m is on the underside, then

~ (0) > 0, ~ (_ mgh ) < 0, ~ (


mgh) < 0,
--;;;: ~(oo) > 0,
+b
al

and equation (23) has two roots. Let ki < k~, then kr is less that the
smaller of these values
mgh mgh
-al+b' a2
338 7. Lagrange's differential equations

and k~ is greater than the larger of them. If h > 0, i.e. the mass m is on
top, equation (23) has real roots under the condition

The expression in brackets should be positive otherwise both roots k~


and k~ are negative. We obtain the following inequality

C31Toi > mgh (val + b+ Va2)2, (7.9.24)

which is the necessary and sufficient conditions to ensure k~ and k~ are


positive. Another derivation of this condition for gyroscopic stabilisation is
suggested in [79].
For positive k~ and k~, the general solution of the system of differential
equations (21) is given by

a D~l) cos (kIt + 'l/JI)


+ D~2) cos (k2t + 'l/J2) ,
(3 D~l) (al +~ k~k+ mgh sin (kIt + 'l/JI) +
3TO I

D I(2) (al + b) k~ + mgh . (k t .1.)


G k sm 2 + 0/2 .
3TO 2

The four integration constants D~1) , D~2) , 'l/JI, 'l/J2 are determined by means
of the initial conditions. The motion is a superposition of harmonic oscilla-
tions with frequencies kl and k 2 . If (for h > 0) inequality (24) does not hold
true, then a and (3 grow without bound as t -+ 00 and use of the linearised
differential equations (21) is meaningless. Analysis of the applicability of
linearising the equations of motion is beyond the scope of this book.
When the rotation axis of the outer ring is vertical we can remove a from
the integrals (19) and (20). Denoting sin{3 = U, we arrive at the differential
equation with the separable variables

(7.9.25)

whose solution is reduced to hyperelliptic quadratures. The case of a bal-


anced gyroscope (h = 0) was studied in detail by Nikolai [70]. Chetaev [21]
indicates how to integrate eq. (25) for h =f. o. Analysis of the stability of
the vertical position {3 = 7f /2 of the rotor, see Fig. 7.2b, is given by Magnus
[61] and Rumyantsev [78].
In the case of the vertical rotation axis of the outer ring, the second
equation in (18) takes the form

aiJ + ~ba2 sin2{3 - C 3TOa cos {3 + mghcos{3 = o. (7.9.26)


7.9 Example of applications of Lagrange's equations 339

The latter equation along with eq. (18) yields a particular solution

(3 =~, /J = 0, a = ao,

describing a uniform rotation of the outer ring and the vertical position of
the rotor axis. Taking
7r
(3=-+c
2
and linearising eq. (26) under the assumption that c is small we obtain

a2E: + (-bo:~ + G3TOO:O - mgh) c = O.


The motion of the rotor axis will be a harmonic oscillation about the ver-
tical position under the following condition

(7.9.27)

The quadratic equation f (ao) = 0 has real-valued roots (0:0)1 and (aO)2 if
(7.9.28)

and they are positive if the angular velocity 0:0 of the outer ring satisfies
the condition

(7.9.29)

These conditions of stability of the vertical position of the rotor axis in


Cardan's suspension were obtained in the above papers by Magnus and
Rumyantsev.

7.9.4 System of two rods


Let us consider a plain system consisting of two rods as shown in Fig. 7.4.
Angles 'PI and 'l/Jl between the rods AlBl and BlGl , respectively, and axis
Alx are taken as the generalised coordinates. The difference 'l/Jl - 'PI is
denoted as (31' The moments of inertia of the rods AlBl and Bl Gl about
axes perpendicular to plane Alxy and passing through point Al and Bl
are denoted by J l and 81, respectively. The lengths of the rods AlBl and
Bl Gl are denoted as Tl and PI' respectively, and the distance Bl G l between
the centre of gravity of the second rod and joint Bl is designated as 81.
The force Fl applied at point Bl is perpendicular to AlBl . The angular
accelerations of the rods are required.
The kinetic energy of the system is

1 2 1 ( 2 2 . .2)
(7.9.30)
T = "2JltPl +"2 mlT l tPl + 2mlT18ltPl'l/Jl cos (31 + 8 l 'l/Jl .
340 7. Lagrange's differential equations

FIGURE 7.4.

The first term is the kinetic energy of rod AIBl rotating about the fixed
axis A. The other terms represent the kinetic energy of rod BI C l calculated
by means of eq. (4.7.7).
The elementary work of the force F I is FI r I 8<pI' that is

U sing the notation

(7.9.31)

we arrive at the equations of motion

(7.9.32)

For the system under consideration, equations in (4.6) have the form

All~l + A12~1 + [1,1; 1] ~~ + 2 [1, 2; 1] ~l~l + [2,2; 1] ~~ = Ql, }


A21 <PI + A 221JlI + [1,1; 2] <PI + 2 [2, 1; 2] <P11JlI + [2,2; 2]1JlI = Q2,
(7.9.33)

indices 1 and 2 being referred to <PI and 1JlI' Comparing eqs. (32) and (33)
we see that the non-zero brackets are

This can be proved by direct calculation using formulae (4.10.3).


7.10 Determination and elimination of constraint multipliers 341

Solving equations (32) for the generalised coordinates yields

1',
7fJ1 =
c 1
~
[A"F,r" m,r,s,,;n~, (": A" + '": A,,) l, }
[A12F1r1 +m1 r 1s 1 sm p1 (4?1All-7fJ1A12)],
(7.9.35)

where ~ = AllA22 - AI2. Clearly, the same expressions are obtained when
the braces are calculated using eq. (4.4). The result obtained is used below
in Sec. 7.11.

7.10 Determination and elirnination of constraint


multipliers
Let us return to Lagrange's equations (1.6)

,\

[s(T)=Qs+LAkaks (s=l, ... ,n), (7.10.1)


k=1

and assume for the sake of notational simplicity that all constraints (holo-
nomic and non-holonomic) are stationary, so that eq. (1.5) is written in the
form
n
L aksqs = 0 (k = 1, ... , l) . (7.10.2)
s=1

Among these equations there can be integrable ones. So, if for a particular
k the following conditions

(r,s = 1, ... ,n)

are satisfied, then the k - th constraint equation can be replaced by a finite


equation

(7.10.3)

where

(7.10.4)

The number of redundant coordinates is equal to the number of integrable


equations in (2).
342 7. Lagrange's differential equations

It is essential to assume that the equalities in (2) are solvable in l of the


n generalised coordinates, that is the deficiency of the matrix

a= (7.10.5)

is equal to zero.
Then one can determine the constraint multipliers from l equations, for
instance the following ones

I
LAkaks = £s (T) - Qs (8 = 1, ... ,l). (7.10.6)
k=1

It follows from this that the square l x l matrix

a' = (7.10.7)
*

is non-degenerating because it is the transpose of the l x l submatrix a* of


matrix (5). Let a qr denote the algebraic adjunct of element aqr of r - th
row and q - th column of the determinant la~1 of matrix (7). Then

I
Ak = I:~I ~akS [£s (T) - QsJ, (7.10.8)

and substituting these values of Ak into the other (n - l) equations in (1)


we find
I
£/+r (T) = Q/+r + L [£s (T) - QsJ Nt+r (r = 1, ... ,n - l) , (7.10.9)
s=1

where
I
N l+r
s 1 ""' ks
= -la'I ~ a ak,l+r· (7.10.10)
* 8=1

The latter sum is the determinant obtained from la~1 by replacing the 8-th
row with the following row

(7.10.11)
7.11 Examples 343

The sum
I
Q: = LAkakS (8= 1, ... ,n) (7.10.12)
k=1

is the generalised constraint fOT"ce corresponding to coordinate qs. In order


to understand the reasoning for this name we mentally eliminate the con-
straints described byeq. (2) and include their reaction forces in the active
forces. Then the 8 - th equation in (1) can be cast in the form

(7.10.13)

On the other hand, as the constraints are ideal they satisfy eq. (6.2.4) in
which variations are related by the equalities
n
Laks8qs = 0, (7.10.14)
s=1

and, by virtue of the theorem of Sec. 6.1, the quantities Q; are represented
by formulae (12).
We conclude now that the generalised constraint forces are defined by
the following equalities

Q; = Es (T) - Qs (8=1, ... ,l), }


I
(7.10.15)
Qi+r = 2: Nt+r [Es (T) - Qs] (T" = 1, ... , n - l) .
s=1

The equations of motion without constraint forces are given by eq. (9).
The constraint equations (2) should be added to them.

7.11 Examples
7.11.1 Four-rod system
Let us consider the four-rod system depicted in Fig. 7.5, [5]. In order to
describe its configuration it is sufficient to introduce the angels 'PI and 'P2'
We also introduce two redundant coordinates, namely angles 'l/Jl and 'l/J2
which are related to 'PI and 'P2 by the following relationships

<PI = T"1 cos 'PI + PI cos'l/J 1 - T"2 cos 'P2 - P2 cos 'l/J2 = 0, }
(7.11.1)
<P2 = T"1 sin 'PI + PI sin'l/Jl - T"2 sin 'P2 - P2 sin'l/J2 - a = 0,

obtained by projecting the pentagon AIBICB2A2Al on the coordinate


axes.
344 7. Lagrange's differential equations

.x c

FIGURE 7.5.

The kinetic energy of the system due to eqs. (9.30) and (9.31) is given
by

T ~2 (A(1)/n 2 + 2A(1) ,;, .i. + A(l) .i. 2 ) +


11 .,.-1 12 .,.-1 If'1 22 If'1

~2 (A(2) . 2
11 if!2
+ 2A(2) . •i. + A(2).i. 2 )
12 if!2 If'2 22 If'2 (7.11.2)

where the subscripts (1) and (2) are respectively referred to the rod pairs
A1Bl, BIG and A 2 B 2 , B 2 G. As F1-LA1B1 and F 2 -LA 2 B 2 , the generalised
forces are equal to

(7.11.3)

In order to avoid repeating the calculations we use equations (9.32) to


write

Adopting the following numbering of coordinates

we can cast matrix (10.5) in the form

P2 sin 'l/J2 -T1 sin if!1 T2 sin if!2 II


- P2 COS'l/J2 T1 cos if!1 -T2 cos if!2 .
7.11 Examples 345

The determinant la~ 1 is given by

l a'*I=I-PI~in'I/JI
P2 sm 'l/J2
Plcos'I/Jl 1 ( )
-P2 COS'I/J2 = PIP2 sin 'l/JI - 'l/J2 .

Using the rule of Sec. 7.10 we obtain

N(l) -rl sin 'PI rl cos 'PI I rl sin ('PI - 'l/J2)


3
1:'*11 P2 sin 'P2 -P2 cos 'P2 = PI sin ('l/JI - 'l/J2)'

I:~II
N(2) -PI sin 'l/JI PI cos'I/Jl I rl sin ('PI - 'l/JI)
3
-rl sin 'PI rl cos 'PI = P2 sin ('l/JI - 'l/J2) .
By analogy
N(l) _ r2 sin ('l/J2 - 'P2) N(2) _ r2 sin ('l/JI - 'P2)
4 - PI sin ('l/J2 - 'l/Jd' 4 - P2 sin ('l/JI - 'l/J2)'

and the equations of motion (10.9) take the form

These should, of course, be considered together with the constraint equa-


tions (1). The quantities on the right hand sides of these equations are the
generalised constraint forces Q~l and Q~2· Also, £1/11 (T) and £1/12 (T) are
the generalised constraint forces Q';pl and Q';p2·
Equations (5) can be obtained easily by means of some very simple rea-
soning. Removing mentally the joint C we replace the action of the rod
system A 2B 2C on AIBI C by the reaction force whose projections on the
corresponding axes are denoted as X and Y. The force in the opposite
direction is applied to the system A 2 B 2 C.
The elementary work of the forces applied to the system AlBIC is

8'W = F l r l 8'P1 + X 8 (rl cos 'PI + PI cos'I/Jl) + Y8 (rl sin 'PI + PI sin 'l/JI)
= rl (FI - X sin 'PI + Y cos 'PI) 8'PI + PI (-X sin 'l/JI + Y cos 'l/JI) 8'I/JI
and the equations of motion become

rl (FI - X sin 'PI + Y cos 'PI) ,


PI (-X sin 'l/JI + Y cos'I/JI)·
By analogy we obtain for the system A 2 B 2 C

r2 (F2 + X sin 'P2 - Y cos 'P2) ,


P2 (X sin 'l/J2 - Y cos'I/J2)·
346 7. Lagrange's differential equations

FIGURE 7.6.

Eliminating the unknown quantities X and Y leads to equations (5). After


Lagrange [55], "an advantage of the previous derivation is that the prob-
lem is reduced to the general formulae yielding all the equations for any
problem".

7.11.2 Plane motion of a heavy rigid body on a string passing


through a fixed ring.
The objective is to derive the equations of motion of a heavy rigid body with
two rings 0 and A, see Fig. 7.6. A flexible inextensible closed string passes
through these rings as well as a fixed ring 0 1 , The mass of the string is
neglected. It is presumed that the centre of gravity C of the body lies in the
plane of triangle 0 1 0 A and the body moves parallel to the vertical plane.
Let the total length of the string be L, the distance OA between the rings
l, the body weight Q, its moment of inertia about the axis perpendicular
to the plane of the motion and passing through point 0 be 8 0 .
The system has two degrees of freedom as the body position can be
determined by the position of the pole 0 of the axes Oxy fixed in the
body. To this end, it is sufficient to know the length 8 = 00 1 and the
angle I.{J shown in Fig. 7.6. Indeed, the lengths of the triangle sides OA = 1
and 01A = L - 1 - 8 are known and we can construct this triangle. This
construction defines the body position and thus the angle 'ljJ. Noticing that
the angle of vertex 0 is equal to 7r /2 - (I.{J + 'ljJ) we have
(L - l - 8)2 = 82 + l2 - 2l8sin(I.{J + 'ljJ)
or
F (8, 'ljJ, I.{J) = 28 [L - 1 - 1sin (I.{J + 'ljJ)]- (L - 2l) L = O. (7.11.6)
This is the constraint equation relating the quantities 8, 'ljJ, I.{J to each other.
It is possible to express, for instance, 8 in terms of 'ljJ and I.{J, however this
7.11 Examples 347

would essentially complicate the expressions for the kinetic and potential
energies and the equations for motion.
The kinetic energy of the rigid body in a planar motion is, due to eq.
(4.7.7),

T ="21 [2
mvo + 8 0 w2
z + 2nw z (VOyXC - VOxYc) 1, (7.11. 7)

where wz = ;p and Xc and yc denote the coordinates of the centre of gravity


of the body referring to axes Oxy.
As the coordinates of the pole 0 are

~o = s cos 'P, "70 = -s sin 'P,


we find

~o = 8 cos 'P - sri; sin 'P, 170 = -8 sin 'P - sri; cos 'P

and then

Vox ~o cos'IjJ + 170 sin 'IjJ = 8 cos ('P + 'IjJ) - Sl{; sin ('P + 'IjJ) ,
VOy -~o sin'IjJ + 170 cos'IjJ = -8sin ('P + 'IjJ) - SI{;COS ('P + 'IjJ).

Expression for the kinetic energy takes the form

(7.11.8)

where, for the sake of brevity

a (0") = Xc sin 0" + yc cos 0", a' (0") = Xc cos 0" - yc sin 0", 0" = 'P + 'IjJ
(7.11.9)

and a" (0") = -a (0"). The potential energy is equal to

II -mg~c = -mg (~o + Xc cos'IjJ - Yc sin'IjJ)


-Q (scos'P + Xc cos'IjJ - Yc sin 'IjJ).

The calculation yields

£8 (T) = m [s - a (0") ~ - a' (0");p2 - S1{;2] ,

£<p (T) = m [s2e;? - sa' (0") ~ + 281{; + s;p2 a (0")] ,


£,p (T) = m [~ ~ - sa (0") - a' (0") se;? - 281{;a' (0") + sa (0") 1{;2]
(7.11.10)
348 7. Lagrange's differential equations

and furthermore

Qs = mgcos'P, Q", = -mg sin 'P, Q'Ij; = -mg (xc sin 'IjJ + Yc cos'IjJ).
(7.11.11)

Differentiating the constraint equation (6) with respect to time yields

8 [L - l - l sin ('P + 'IjJ) 1- sl (p + ~ ) cos ('P + 'IjJ) = o.


Using eq. (6) the above equation becomes

8 (L - 2l) L ( .)
2s2 l - <P + 'IjJ cos ('P + 'IjJ) = o. (7.11.12)

Differentiating again with respect to time yields

82 ) (L - 2l) L ( .. )
S
( 2s2 - s3 l - ~ + 'IjJ cos ('P + 'IjJ) + ( .)2
<P + 'IjJ sin ('P + 'IjJ) = o.
(7.11.13)

From eq. (12) it follows that

0'IjJ (L-2l)L 0'IjJ


(7.11.14)
as 0'P = -1.

By virtue of eq. (10.9), considering 'IjJ as a redundant variable, we obtain

(7.11.15)

These equations of motion should be considered together with eqs. (6),


(12) and (13). We notice that

Q'Ij; - Q", = -mg (xc sin 'IjJ + Yc cos'IjJ - ssin'P) = -mgT)c = m?',
(7.11.16)

that is, this difference is equal to the moment of the weight about axis 0 1 (.
The latter equation in (15) thus expresses the angular momentum theorem

(7.11.7)

where kf'
stands for the projection of the angular momentum about point
o on axis 0 1 (. Due to (4.8.10) it is equal to
(7.11.18)
7.12 Generalised reaction forces of removed constraints 349

Here, by virtue of eq. (4.8.9), the projections of the angular momentum


vector Q are as follows

Qf. =m (VOf. - ;Prcry) , Qry =m (VOry + ;Prcf.) , (7.11.19)

where rCf. and rCry denote the projections of vector rc = DC on axes Ol~'T/
rCf. = Xc cos'ljJ - Yc sin ,¢, rCry = Xc sin'¢ + Yc cos '¢. (7.11.20)

Applying now the above formulae for projections of the pole velocity and
using notation (11) we obtain

K?l = m [_82(P + 8;Pa' (a) - sa (a) - s(pa' (a) + ~;P] . (7.11.21)

Having calculated the time-derivative we arrive at relationship (17) in


which e1/1 and ecp are defined by eq. (10). This serves as a control for the
calculation performed.
The differential equations (15) admit one first integral which is the inte-
gral of energy

T+II = h,
expressions for T and II having been derived earlier.

7.12 Generalised reaction forces of removed


constraints
Let us consider a system with n degrees of freedom subject to holonomic
constraints. Its configuration is specified by the generalised coordinates
ql, ... , qn' We remove mentally a number of the constraints whose reaction
forces are required. The number of degrees of freedom increases then up to
n + m and, in order to describe the configuration, we need an additional m
generalised coordinates qn+I, ... , qn+m' Let us choose these additional coor-
dinates so that we return to the original system when all qn+I, ... , qn+m are
equal to zero. In other words, motion of the system is considered with re-
dundant coordinates qn+I, ... , qn+m, the constraint equations (1.4.8) having
the maximum simple form

Fn+1 = qn+l = 0, ... ,Fn+m = qn+m = O. (7.12.1)

This approach was applied in Sec. 6.5 while considering static problems.
Using eq. (10.4) we have

oFn+r = on+r,s
an+r,s = -;:,}-- (r = 1, ... ,m; 8 = 1, ... ,n + m ) (7.12.2)
uqs
350 7. Lagrange's differential equations

and by virtue of eq. (10.12) the generalised constraint forces are as follows

Q: =
m
L .An+ran+r,s =
r=l
° (8=1, ... ,n), }
(7.12.3)
Q~+l = .An+l (l = 1, ... ,m).

The equations of motion (10.1) take the form

£s (T) = Qs (8 = 1, ... ,n); } (7.12.4)


£n+r (T) = Qn+r + .An+r (r = 1, ... ,m).

Here the kinetic energy and the generalised forces are constructed under
the assumption that the constraints given by eq. (1) are absent. When
equations (4) are constructed they should be considered together with the
constraint equations. This means that in eq. (4) we should put

qn+r = 0, rin+r = 0, iin+r = 0, (r = 1, ... ,m). (7.12.5)

In the case of stationary constraints the equations of motion in the ex-


plicit form have the form of eq. (4.6), i.e.

n+m n+mn+m
2: Asuiiu + 2: 2: [a, p; 8] riurip = Qs, (8 = 1, ... ,n), (7.12.6)
u=l u=l p=l

n+m n+mn+m
2: An+r,uiiu + 2: 2: [a, p; n + r] rio-rip = Qn+r + .An+ro (r = 1, ... ,m).
0-=1 u=l p=l
(7.12.7)

Consideration of the non-stationary constraints causes no complications


except for cumbersome expressions. Now accounting for eq. (5) we obtain
instead of (6) and (7)
n n n
2: A~uiiu + 2: 2: [a, p; 8]D riurip = Q~ (8 = 1, ... ,n), (7.12.8)
u=l u=lp=l

n n n
2: A~+r,uiiu + 2: 2: [a, p; n + r]D riurip = Q~+r + .An+r , (r = 1, ... ,m).
u=l u=lp=l
(7.12.9)

The zero superscript is used here to denote the values of parameters when
the redundant coordinates equal zero.
7.12 Generalised reaction forces of removed constraints 351

Equations (6) and (7) were obtained by means of the kinetic energy in
the form

1 n+mn+m 1 n n
T 2" L L Asaqsqa = "2 L L Asaqsqa (7.12.10)
s=l a=l s=l a=l
m n 1 m m

+L L An+r,aqn+rqa + 2" L L An+r,n+tqn+rqn+t,


r=la=l r=lt=l

where coefficients Asa are functions of all variables q1, ... , qn+m' The gen-
eralised forces in these equations are calculated by using expressions for
the elementary work of the active forces due to the virtual displacements
defined by the variations t5q1, ... , t5qn+m
n m

t5'W = L Q st5qs + L Qn+rt5qn+r. (7.12.11)


s=l r=l

The kinetic energy TO of the original system is obtained from expression


(10) at qn+r = 0 and qn+r = 0 (r = 1, ... , m)

(7.12.12)

Assuming qn+r = 0 in the expressions for the generalised forces we find


their values in eqs. (9) and (8). Q1, ... , Qn are equal to the generalised
forces in the equations of motion of the original system. Using eq. (4.6)
and the definition of the square brackets, these equations have the form

The square brackets in eq. (8) are given by

where zero indicates that qn+1, ... , qn+m are set to zero after differentiation
whereas those quantities in parentheses are taken to be zero before differen-
tiation. However it makes no difference since the derivatives are taken with
respect to variables qs for which s = 1, ... , n while the other variables are
set to zero. Hence, equations of motion (8) are identical to the equations
of motion of the original system (13).
352 7. Lagrange's differential equations

Let us notice, that due to the above reason, the brackets in eq. (9) are
determined as follows

[a p-n+r jo = -1 (OAa ' n+r + oAp 'n+r -oqn+r


oAap)O
--
, , 2 oqp oqa

= ~ [OA~,n+r + oA~,n+r _ (OAa p )0]. (7.12.14)


2 oqp oqa oqn+r

In order to determine them it is necessary to know the values of coefficients


Aap in expression (10) for the kinetic energy with accuracy up to the terms
linear in qn+r' The values of An+r,a should be known only at qn+r = 0,
that is only A~+r,a are needed and the coefficients An+r,n+t (r, t = 1, ... , m)
are not needed at all.
The generalised accelerations can be replaced by their expressions due
to eq. (8). According to eq. (4.7) we have

(7.12.15)

where the braces are calculated for the matrix of the coefficients of the
°
quadratic form (12) and (A;;}) denotes the elements of the inverse matrix.
Substitution into eq. (9) for the generalised reaction forces of the removed
constraints yields

An+r - ~ A~+r,a ~ ~ {;} °qpq>. + ~ A~+r,a ~ (A;;}) °Q~ +


n n
L L [a, p; n + rjO qaqp - Q~+r' (7.12.16)
a=l p=l
We introduce the notation
n
LA~+r,a(A;;})o=M~+r (r=l, ... ,m;p=l, ... ,n). (7.12.17)
a=l
These values are calculated as follows: in the determinant of the quadratic
form

A~l A~2 A~n


Ag 1 Ag2 Agn
IAol = (7.12.18)
A~l A~2 A~n
the p - th row is replaced by the following row

(7.12.19)
7.13 Geometrical interpretation of the generalised constraint forces 353

Dividing the result by IAol yields M~+r'


Recalling values (4.4) of the braces we have, due to eq. (17),

~A~+r'O"{:.Ar = ~tA~+r,O" (A;;;)o [p,.A;TjO = tM~+r [p,.A;TjO.


(7.12.20)

Substituting these into expressions (16) we arrive, after some changes of


indices, to the equalities

.An+r ~ ~ qaqp {[a, p; n+ rjO - t M~+r [p,.A; TjO } +


[t M~+rQ~ - Q~+r1 (r = 1, ... , m). (7.12.21)

These equalities are our aim. They give expressions for the generalised
reaction forces of the removed constraints in terms of the prescribed forces
and the generalised velocities. The generalised reaction forces are quadratic
forms of these velocities.
Each of the quantities .An+r is determined by eq. (21) independently of
the others. Thus we can find .An+l, ... , .An+m separately by introducing the
redundant coordinates one by one.

7.13 Geometrical interpretation of the generalised


constraint forces
An example of the fruitful application of the Riemannian geometry to dy-
namics is the problem of the constraint forces considered in the previous
section.
The basic ideas of Riemannian geometry are briefly outlined in Sec. B.13.
As explained in Sec. 7.8 a Riemannian manifold Rn with the quadratic form
TO (dt)2 corresponds to the system of material points under consideration.
The covariant components of the metric tensor on this manifold are denoted
as aOl.{3' The manifold Rn is a part of manifold Rn+m with the quadratic
form T(dt)2. This means that the points of Rn+m with the following pre-
scribed values of the generalised coordinates

qn+ 1 ° 1 ,... ,qn+m = qn+m


= qn+ ° (7.13.1)

belong to Rn. In other words, the considered system of particles can be ob-
tained from the system corresponding to Rn+m by imposing m constraints
described by relationships (1). Setting dqn+s = 0 in the expression for the
354 7. Lagrange's differential equations

kinetic energy of the system Rn+m

T (dt)2 = ~ (9o!3dq Odq{J + 29o,n+sdqodqn+s + 9n+s,n+kdqn+sdqnH)


(7.13.2)

(summation over Greek indices from 1 to n and over Latin indices from 1
to m is assumed) we arrive at TO which is the kinetic energy of the system.
Therefore

(7.13.3)

The force corresponding to the representative point of the manifold Rn+m


is given by the vector Q+A with the covariant components Qo and Qn+s +
A n +s . The elementary work of the given forces in the virtual displacement
of the particles of the system Rn released from the constraints is

(7.13.4)

whereas

(7.13.5)

implies the elementary work of reaction forces of the released constraints.


Newton's second law, for Rn+m, is written in the form

w=Q+A (7.13.6)

and the problem reduces to determining the acceleration vector w at points


of Rn. The velocity vector in Rn+m is given by

Its time-derivative is

However at points of Rn+m belonging to Rn we have, due to eq. (1),


qn+s = 0 ijn+s = O. (7.13.7)
Thus,

(v)O = Poijo + (ro!3)qOq!3,


where Po = (ro)o denotes the basis vectors of Rn. Using the relationship
(B.13.21), we replace (r o!3)o by its value and arrive at the equality

wo (v)o = P, (ij' + {a~} aqoq!3) +


([a, {3, n + tlo - M;{+t [a, (3, I'la) qOq!3r~+t. (7.13.8)
7.13 Geometrical interpretation of the generalised constraint forces 355

The index a denotes the quantities calculated by means of the metric


Rn. Thus, according to eqs. (8.7) and (8.8) the first group of terms in (8)
represents the acceleration vector (w) a in Rn

(7.13.9)

The second group determines the vector


- -_·a·f3
Wo q q ([ a" (3 n + tl 0 - M"!n+t [a, (3.,"f1a) ron+t . (7.13.10)

Since

r~+t. p,,! = 0,

vector Wo is orthogonal to the space Rn. Vector Wo in eq. (8) means ac-
celeration at points Rn+m belonging to Rn. It is equal to the geometric
sum of the acceleration vector in Rn and vector Wo considered at points
of Rn. Relationship (8) generalises formula (6.3) for motion of the particle
on a surface. It is pertinent to note that calculation of the coefficients of
the second quadratic form of the surface ba f3 is performed according to the
rule of eq. (10), see also the end of Sec. B.13.
Turning now our attention to eq. (B.13.28), we represent the force vector
Q + A at points Rn in the form

The first component is calculated in metric Rn and is equal to aaf3 (Q)f3 Pa,
i.e. it does not contain the components of vector A. Newton's second law
in the extended form, eq. (6), takes the form

Therefore, we obtain equations of motion in Rn in the form (8.12)

and the expressions for the covariant components of vector of the constraint
force

(An+t)o = qaqf3 ([a, (3; n + tlo - MJ+t [a, (3; "fla)


- [(Qn+t)o - M::+t (Qa)a] (t = 1, ... ,m)(7.13.13)
which is orthogonal to the space Rn. Taking into account the changed no-
tation we arrived at expressions (12.21). This new result is instructive as
356 7. Lagrange's differential equations

...--- ---v

FIGURE 7.7.

it links the laws of mechanics and geometric images. The forces of ideal
constraints are orthogonal to the space in which the representative point
moves. Only that part of vector v which belongs to Rn is needed for spec-
ifying the motion. Removing constraints we find the other components of
this vector and the complete meaning of the vector v becomes clear when
all constraints are removed, i.e. in the space E 3N .

7.14 Application to planar systems of rods


1.14.1 Physical pendulum
We consider a physical pendulum which is a heavy rigid body of weight G
having a vertical plane of symmetry. The body can rotate about a fixed
axis 0 which is perpendicular to this plane, see Fig. 7.7. We mentally cut
the pendulum into the bodies 1 and 2 using cross-section SKS. We search
the resultant force and the resultant moment of the reaction forces in this
cross-section, i.e. the forces which ensure the integrity of the pendulum.
The kinetic energy of the pendulum and the potential energy of the
weight are as follows

(7.14.1 )

where eo denotes the moment of inertia of the pendulum about the rota-
tion axis and OC = c stands for the distance between the centre of gravity
C and the rotation axis.
After the cutting, the body 2 possesses three degrees of freedom. We
take the displacements q2 and q3 as well as the angle ql + q4 , see Fig. 7.7,
as the parameters describing the position of the system. In the original
°
configuration q2 = 0, q3 = 0, q4 = which corresponds to conditions (12.1)
of the redundant coordinates.
7.14 Application to planar systems of rods 357

For the sake of simplicity we assume that the centres of gravity C l and
C 2 of bodies 1 and 2 lie on a straight line OC passing through point 0, the
centre of gravity of the pendulum C also lying on the line OC.
The kinetic energy of the system released from the constraints, that is
pendulum 1 and body 2, is given by the expression

(7.14.2)

where 8~ denotes the moment of inertia of the pendulum 1 about the


rotation axis 0, 8c2 denotes the moment of inertia of the body 2 about
the parallel axis passing through the centre of inertia C2 of this body, m2
is its mass and VC2 denotes the velocity of point C 2 .
The coordinates of the latter point are

XC2 a cos ql - q2 sin ql + q3 cos ql + C2 cos (ql + q4) ,


YC 2 a sin ql + q2 cos ql + q3 sin ql + C2 sin (ql + q4) .

Thus,

vb2 xb2 + iJb2 = [(a + C2) 2 + 2 (a + C2) q3] qi +


2C2 (a + C2) qlq4 + 2 (a + C2) (iIq2 + ... ,

where dots imply terms which are out of interest as they contain the terms
that are of order higher than the second. We obtain
1
T = 2 {[80 + 2m2 (a + C2) q3] iIi + 2A~2qlq2 + 2A~4qlq4 + ... } .
(7.14.3)
Here

denotes the moment of inertia of the pendulum about the rotations axis 0
and

The potential energy of the system is given by

II = -GlXC1 - G2XC2 = -GlCl cosql -


G2 [(a + q3) cos ql - q2 sin ql + C2 cos (ql + q4)]
-GcCOSql - G2q3COSql + G 2 (q2 + c2q4) sinql+,.·· ,

where
358 7. Lagrange's differential equations

Thus
Q~ = -Gcsinql, Qg = -G2 sinql, Qg = G 2 cosql, Q~ = -G2c2sinql·
By formulae (12.17) we have

M 2l _ m2 (a + C2) 1
- , M3 =0,
80
In addition to this
[1,1; 3]° = -m2 (a + C2),
the other brackets being identically equal to zero. Calculation due to eq.
(12.21) leads to the following expressions

.
G sm [G2 m2 (a + C2) C2]
A2 ql (j - 80 '

A3 - [m2 (a + C2) q~ + G2 cos ql] ,


G2
[C2(j m2 C2 (a + C2) + 8 C 2 ] G .
- C 80 smql·

In this example A2 and A3 are the transverse and axial forces, respec-
tively, and A4 the bending moment in the cross-section SKS. They are the
generalised forces corresponding to the generalised coordinates q2, q3, q4.
The positive directions of the forces and the moment acting on the cut
part coincide with the positive directions of the generalised coordinates.

7.14.2 Generalised constraint forces in plane mechanisms


We begin by considering an open chain of n elements with n degrees of
freedom. Figure 7.8 shows a chain of the members in series, however the
forthcoming derivation will be valid for chains with branch members. A
chain member is conditionally shown by a line segment linking two succes-
sive joints. However it does not lead to loss of generality as the centres of
gravity are not assumed to lie on these lines. The member with the fixed
axis of rotation is taken as the driving member and has the index 1. The
member between joints Ai and Ai+l has index i. The point Ai is adopted
as the origin of the coordinate system for the i - th rod, the axes being
given by the unit vectors ei and e~. Vector ei is directed from Ai to Ai+l
and has angle 'Pi with the fixed axis A1x. The position of the centre of
gravity C i in this coordinate system is described by the position vector

The velocity vectors of the joints are related to each other by means of the
following recurrent equalities
(i = 2, ... ,n), (7.14.4)
7.14 Application to planar systems of rods 359

FIGURE 7.8.

where li-l denotes the length of A i - 1 A i . The kinetic energy of the i - th


member is calculated by eq. (4.7.7) as

2Ti miv; + 2mivi . (k<Pi x r~) + 8 i <P;


miv; + 2mi<pi (slvi· e~ - S~Vi· ei) + 8 i <P;, (7.14.5)

where mi and 8 i denote the member mass and its moment of inertia about
the axis of the joint Ai , respectively. The unit vector perpendicular to the
plane of the chain is denoted by k, so that k<Pi is the angular velocity vector
of the i - th member.
Thus, we have

T 1 .2
1 = 28-1'Pl'
18 . 2 1 l2 . 2 l . . ((2)
rp
12 = 2-2'P2 + 2m2 (3 (2) . (3 )
l'Pl +m2 1'P2'PI SI cos 21 - 8 2 sm 21 ,

r.3 = 21 8- 3'P3
. 2 + 21 m3 (l21 'PI
. 2 + l22'P2
. 2 + 2l 1 l 2'Pl
. .'P2 cos (3)
21
+ m3'P3
. x
[h <PI (8~3) cos (331 - S~3) sin (331) + b<P2 (s~3) cos (332 - 8~3) sin (332) ]

etc., where the difference of angles is introduced as follows

(7.14.6)
360 7. Lagrange's differential equations

Calculating the sum of expressions (5) we obtain the kinetic energy of


the chain in the form of a homogeneous quadratic form
n n n
2T = Al1<P~ + 2 LAlS<Pl<Ps + LLAks<Pk<Ps, (7.14.7)
s=2 s=2k=2
where the coefficients Akk are constant and Aks depend only on (3 sk' so
that
_ ,
BAsk - 0 r
.../..
I S,
k; (7.14.8)
B
'Pr
The elementary work of active forces acting on points of the i - th member
is reduced to the form of eq. (5.25)

(7.14.9)

where Vi and m~ denote the resultant force and the resultant moment of
these forces about the axis of joint A, respectively, and Ori is the virtual
displacement of the joint Ai, for which the recurrent relationship
(7.14.10)

holds. The elementary work of the active forces for the whole chain is
n n
O'W = L Qso'Ps = QlO'Pl + L Qso'Ps· (7.14.11)
s=l s=2
The chain of n members becomes a mechanism under n - 1 constraint.
Their equations

(7.14.12)

can be assumed to be resolved in 'P2'· .. ,'Pn


(7.14.13)
Then
(7.14.14)

and the expression for the kinetic energy (7) can be written in the form

T o -_ "218* ('PI ).2


'Pl· (7.14.15)

Here
n n n
8* ('PI) = A~l + 2 L A~sf; + L L A~sfkf; (7.14.16)
s=2 k=2s=2
7.14 Application to planar systems of rods 361

is referred to the moment of inertia of the mechanism with respect to axis


Al of the driving member. Zeros indicate that the quantities are calculated
for the values of <P2, ... , <Pn defined by the constraint equations.
The elementary work of the active forces (11) becomes

(7.14.17)

m z being the torque for the driving member. The equation of motion for
the mechanism can now be written down as follows

co
VI
(TO) = 8* ( ).. 1 d8* . 2
- <PI <PI + "2 d<P1 <PI = m z · (7.14.18)

Let us mentally remove the constraints, then instead of relationships (13)


and (14) we can take

(7.14.19)

where it is sufficient to retain the terms the order not higher than the first
in qs and qs.
We begin with the kinetic energy. In order to obtain its expression we
should replace in eq. (7) the quantities <Ps and rps by their expression from
eq. (19). The result is

8A1S)O qr1f'·s<P1 + AO··


Als<P1..<Ps = [AOIs + u~ (a- 2
lsqs<P1,
r=2 <Pr

.. = [AOks + ~
Aks<Pk<Ps u (8A
kS )O qr1f'f,·2
-8-- k s<P1 + AOks (f'·
kqs + f'·)·
sqk <PI'
r=2 <Pr
(k, s = 2, ... , n),
(7.14.20)

where, in accordance with eq. (8),

~ (8AkS)O qr =
~ -8-
(8AkS)O
8{3
(qs - qk ), ~ (8A1S)O qr =
~ a-
( 8A
8{3
1S)O qs·
r=2 <Pr sk r=2 <Pr sl
(7.14.21)

Due to eqs. (7) and (16) the kinetic energy of the mechanism released from
the constraints is given up to the above accuracy by the equality

2T =

(7.14.22)
362 7. Lagrange's differential equations

While deriving this equation we noted that


n n n n
L L A~8 (J~q8 + I~qk) = 2 L L A~sf~q8
8=2 k=2 8=2 k=2
and

Expression (11) for the elementary work should be taken as


n n
8'W = Q~8<p1 +L Q~ (J~8<p1 + 8q8) = m Z 8<p1 +L Q~8q8. (7.14.23)
8=2 8=2

Now we have all of the equations and we can apply formula (12.21). In our
case n = 1 and it takes the form

Ar = <pi ([1,1; rjO - M; [I, 1; IjO) + M~mz - Q~ (r = 2, ... ,n) .


(7.14.24)

The coefficients of the kinetic energy are reduced to the form

n
At + L 8Ak8/~·
k=2
Therefore,

[1,1; rjO dA~r


- ~ (dA~r
- + L...-
d<p1
- - I'k + AOkr k -
k=2 d<p1
1")
( ~A(31r)0
81
I~ - t (~A(3kr)O I~/~.
k=2 rk
Taking into account that

we have
7.14 Application to planar systems of rods 363

FIGURE 7.9.

and

[1,1; 11° ~ ~ ~~:, M; ~ e. ~<Pl) (Aj, + t, A2,j;) .


The generalised force of the r - th constraint is equal to

Ar = -~i [ (~~:~ ) + ~ { (~~:: ) 1~2 - A~rlr } +


0

"21 de*
d'Pl
(0 ~ 0 ') 1 1 m (0 ~ 0 ') 0
AIr + 6 Akrlk e* ('PI) + e* ('PI) AIr + 6 Akrlk -Qr·
z

(7.14.25)

7.14.3 Crankshaft mechanism


This mechanism is shown in Fig. 7.9 and can be obtained from a three-
member chain by imposing the following constraints

sin'P2 = -{.L sin 'PI' 'P3 =0 ( ,....II. -- _ll2I) . (7.14.26)

It follows from these equations that


' = _/leos 'PI , i"2,....,....
I 2,.... = (1 _ /12) II. sin3'PI, 13' = iJ" = O. (7.14.27)
cos 'P2 cos 'P2
The above expressions for T I , T 2 , T3 yield the following coefficients
All e I + (m2 + m3) li, A22 = e 2 + m3l~, A33 = e3 ,
AI2 m2h (si 2) cos{321 - S~2) sin{32I) + m3hl2 cos {32I '
A I3 m3h (si3) cos (33I - s~3) sin (33I) ,
A 23 m3l2 (si3) cos (332 - s~3) sin (332) .
364 7. Lagrange's differential equations

The reduced moment of inertia and its derivative with respect to <P1 are

8* (<P1)
d8* (<P1)
d<P1

and the generalised constraint forces due to (25) take the form

The generalised constraint force A2 represents the moment of the reac-


tion force applied to the connecting rod under the condition <P3 = o. The
moment of the reaction force ensuring no rotation of the slide block under
the first constraint in eq. (26) is denoted as A3. The problems of this sort
are comprehensively studied in [38].
It is easy to obtain the moment of the reaction forces A2 by using the
motion of the two-member chain (9.32). The second of these equalities
should be written in the form
·· dA 12 ·2 Q \
A 12<P1 + A··
22<P2 - d(321 <P1 = 2 + "2,

since the partial derivative with respect to <P1 can be replaced by the deriva-
tive with respect to -(321. Substituting

<P2 = f~<P1' <P2 = f~' <p~ + f~<P1


and replacing <P1 by its value due to the equation of motion (18) of the
mechanism we obtain the required value of A2.

7.14.4 System of two rods


The reaction force in joint A2 of the two-rod system depicted in Fig. 7.10
is required. The system is loaded by force F at point A 3 .
Separating the second rod at joint A2 we denote the projections of the
imaginary displacement of this point on axes x and y as q3 and Q4, respec-
tively. The velocity of this point is then

V2 = hq1e~ + q3 i 1 + Q4 i 2,
7.14 Application to planar systems of rods 365

FIGURE 7.10.

where hand h are the unit vectors of axes x and y, respectively. The
kinetic energy of the first rod is
1 9 .2
T 1 = "2-1q1·

For calculation of the kinetic energy of the second rod it is sufficient to


retain terms of order not higher than the second in ti3 and ti4 as the higher
order terms are of no concern for calculating the generalised constraint
forces. Then we have

T = "21 {9- 2q2·2 + m2 (l21 q1·2 + 2l· · cos q1


1 q1 q4 .q3
- 2l 1 q1 ..sm q1 ) +
2m2sti2 [h til cos (q2 - q1) - ti3 sin q2 + ti4 cos q2]} ,
where s denotes the distance between the centre of gravity and the joint
A 2 . Now we have

T = T o + 2Ao13q1q3
.. + 2Ao· · + 2Ao··
14q1q4 23q2q3 + 2Ao··
24q2q4, (7.14.29)

where TO denotes the kinetic energy of the original two-rod system

T o = "21 (A llq1
·2 ..
+ 2A 12q1q2 + A 22q22. ) , (7.14.30)

where the coefficients do not differ from those at q3 = q4 = 0

All = 9 1 + m2l~, A22 = 92, A12 = m2sh cos (q2 - qd


The other coefficients are equal to

A~3 = -m2h sin q1, Ag3 = -m2S sin q2 ,


A~4 = m2h cos q1 , Ag4 = m2s cos q2.
366 7. Lagrange's differential equations

The elementary work of force F is given by

D'W = F· 15 (he~ + i 1q3 + i 2q4 + l2ei)


= F· (he~Dq1 + i 1Dq3 + hDq4 + l2e§Dq2)
= (-X sinq1 + Y COSq1) Dq1 + XDq3 + YDq4 + (-X sinq2 + Y COSq2) Dq2
and the generalised forces are

Q1 = -Xsinq1 + YCOSq1, Q3 = X,
Q2 = -X sinq2 + Y COSq2, Q4 = y.
Formulae (12.21) for n = 2 take the form

AT = qi M; [1,1; 1]° - M; [1,1; 2]°) +


([1,1; r]o - (7.14.31)

2q1q2 ([1,2; r]o - M; [1,2; 1]° - M; [1,2; 2]°) +


q~ ([2,2; r]o - M; [2,2; 1]° - M; [2,2; 2]°) + M;Q~ + M;Qg + Q~,
where r = 3,4. In our case the non-zero brackets are

[2,2; 1]° = -m2hs sin (q2 - q1), [1,1; 2]° = m2hs sin (q2 - q1),
[1,1;3]° = -m2hcosq1, [2,2; 3]° = -m2S cos q2,
[1,1;4]° = -m2h sinq1, [2,2; 4]° = -m2S sin q2.

We also have

and moreover, according to eq. (12.17),

M1 = _ m2 I h sin ql s sinQ2 I 1112 I


3 11101 1112 1122 ' ssinQ2 '

I h cos q1 I
Ml = m2
4 11101 1112
S cosQ2
1122 ' S~~2Q21
What remains is to substitute the formulae obtained into eq. (31). We
restrict our consideration to the value of A3 under the assumption that
force F is absent. We obtain

A3 =- T~~I {11ll [1122 COSQl - m2 s2 sinQ2 sin (Q2 - Ql)] -

22 2 ( }.2 m2 s
m2 l 1 s cos Q2 -Q1)COSQ2 Q1 -11101 X

{1122 [11ll cos Q2 + m2li sin Q1 sin (Q2 - Q1)] - m~li s2 cos (Q2 - Qd cos Q1} q~.
7.15 Cyclic coordinates 367

The same expression for >'3 could be obtained by constructing equations of


motion for the centre of inertia of the rod A2A3

where iiI and ii2 should be replaced by their values due to the equations
of motion of the two-rod system, i.e. byeq. (9.35). This solution would be
shorter however the automation of the calculation would be lost.

7.15 Cyclic coordinates


Let us agree to refer to a generalised coordinate qs of a holonomic system
as being cyclic if the following conditions are met: the corresponding gen-
eralised force Q s is equal to zero whilst the other generalised forces as well
as the kinetic energy (i.e. To and coefficients Ask and Bsk) do not explic-
itly contain this coordinate. The generalised velocity corresponding to the
cyclic coordinate is called the cyclic generalised velocity. The coordinates
which are not cyclic are called the positional coordinates. If the forces are
potential forces, then, by the above definition, those coordinates which do
not appear in the expression for the kinetic potential are cyclic forces.
Let the system have n degrees of freedom and its configuration be de-
scribed by the generalised coordinates ql, ... ,qn, among them there are m
positional coordinates ql, . .. ,qm and n-m cyclic coordinates qm+ 1, . .. ,qn' .
Then,

Qm+l = 0, ... ,Qn = 0; Qs = Qs (ql,'" ,qm) (s = 1, ... ,m) (7.15.1)

and

(7.15.2)

and in the case of potential forces

(7.15.3)

The expression for the kinetic energy can be split into the components

T = T2 +Tl +To = T; + U +T2* +T~ +T~* +TO(ql,'" ,qm;t).


(7.15.4)
368 7. Lagrange's differential equations

Here Ti. and Ti.* are the quadratic forms of the positional and cyclic coordi-
nates, respectively, and U is a bilinear form of both types of the coordinates
1 m m

Ti. = 2L L A sk q8qk,
k=18=1
1 n-mn-m
Ti.* = 2L L A m+8,m+kQm+8Qm+k, (7.15.5)
k=1 8=1
m n-m
U= LL A 8,m+kQ8Qm+k'
8=1 k=1
The linear part of the kinetic energy can also be represented by the two
components
m n-m
T! = LB8Q8'
8=1
T!* = L
8=1
B m+8Qm+8' (7.15.6)

The quadratic forms Ti. and Ti.* are obtained by means of the quadratic
matrices
All AIm A m+1,m+l A m+1,n
A* = A** =
AmI Amm A n ,m+l Ann
(7.15.7)

These matrices are non-degenerate as their determinants IA * I and IA ** I are


the diagonal minor determinants of IAI which are positive by virtue of the
Sylvester inequalities (4.1.10).
We turn now to the differential equations of motion. Noticing that

~=o (8
!} = 1, ... , n - m ) , (7.15.8)
uqm+8
and taking into account eq. (1), we split them into two groups: equations
of motion for the positional coordinates

!!:... aT _ aT
n
= Q8
(8=1, ... ,m) (7.15.9)
dt aQ8 uq8
and those for the cyclic coordinates
d aT
---=0 (8=1, ... ,n-m). (7.15.10)
dt aQm+8
The latter equations can be integrated immediately to obtain the n - m
first integrals of the equations of motion
aT
-!}-.-- = Pm+8 = i3 m +8 (8 = 1, ... , n - m) (7.15.11)
uqm+8
7.15 Cyclic coordinates 369

expressing that the generalised cyclic momenta are constant and denoted
by f3m+s. An expanded form of the latter equation is
oT:;* oU oTi*
- 0 · + -0-·- + -0-·- =f3 m +s (s = 1, ... ,n-m) (7.15.12)
qm+s qm+s qm+s
or
n-m m
L Am+s,m+ktlm+k = f3m+s - L Am+s,ktlk - Bm+s (s = 1, ... ,n - m) .
k=l k=l
(7.15.13)
This system of linear equations admits a solution in generalised cyclic ve-
locities as its determinant IA**I is positive. These generalised cyclic veloc-
ities depend on the positional generalised coordinates and velocities, the
constant momenta and time

tlm+r = f m+r (q!, ... ,qm, tl1' ... ,tlm; f3m+ 1, ... ,f3n ; t) (r = 1, ... ,n - m) .
(7.15.14)

These forms are linear in f3m+s - Bm+s and tlk. Indeed, let
Am+1,m+1
(7.15.15)

be the matrix inverse to A **, i.e.

Am+s,m+k _ _1_~
- IA**I m+k,m+s, (7.15.16)

~m+k,m+s being the algebraic adjunct of the element Am+k,m+s of the


determinant IA ** I, then
n-m m n-m
tlm+r = L (f3 m+s - Bm+s) Am+r,m+s - Ltlk L Am+s,kAm+r,m+s.
s=l k=l s=l

l
(7.15.17)
The quantities

N m+r = ~ (f3m+s - Bm+s) Am+r,m+s,


n-m (7.15.18)
M!.+r = L Am+s,kAm+r,m+s
s=l
are quotients with IA**I in the denominator. The numerators are obtained
from the determinant IA**I in which the elements of the r - th row
370 7. Lagrange's differential equations

are replaced by

and respectively by

We then obtain
m
qm+r = Nm+r - L M!+rqk (r = 1, ... , n - m). (7.15.19)
k=l

These are the expanded expressions for functions (14). When we substitute
them into the equations of motion (9)

(s=l, ... ,m), (7.15.20)

the positional generalised velocities and coordinates, time t and the con-
stant cyclic momenta appear in these equations.

7.16 The Routhian function


A new form of the differential equations of motion was suggested by Routh.
The idea is to introduce, instead of the kinetic energy T, another function
R depending upon the positional coordinates and velocities, time and the
cyclic constant momenta. This function is defined as follows

(7.16.1)

Let us construct the expressions for the derivatives

oR oR oR
(k = 1, ... , n - m; r = 1, ... , n - m),
oqk' Oqk' of3m+r

taking into account that the arguments qk, qk, f3 m+r appear on the right
hand side of definition (1) both explicitly and in terms of functions fm+r.
Then we have
7.16 The Routhian function 371

since, byeqs. (15.11) and (15.14),


aT
- 8 1 =/3m+r
m+r
and the sums in eq. (2) cancel out. This calculation is also valid for the
derivatives with respect to qk
8T
(k= 1, ... ,m). (7.16.3)
8qk
By analogy we find

and thus

qm+s = - 8/38R (s = 1, ... , n - m) . (7.16.4)


m+s
Inserting expressions for the derivatives (2) and (3) into the differential
equations of motion for the positional coordinates (15.9) we obtain

(7.16.5)

These equations suggested by Routh have the structure of Lagrange's equa-


tions, the part of the kinetic energy being played by the Routhian function
R. These equations contain only the positional coordinates as well as the
generalised velocities and accelerations corresponding to these coordinates.
For this reason, this approach is referred to as the method of ignoration of
cyclic coordinates whilst these coordinates are referred to as being ignor-
able or hidden. In contrast to these the positional coordinates are termed
explicit.
By integrating the system of m differential equations of the second order
we determine the generalised coordinates and velocities as functions of time
t, n - m constant values /3m+s and the 2m integration constants

qs = qs (t; /3m+l'··· , /3 n , Gl ,··· , G2m ) ,


(7.16.6)
qs =qs (t;/3m+l'··· ,/3n,Gl , ... ,G2m ) (s=I, ... ,m). }

These expressions should be substituted into equalities (4). Then determi-


nation of the cyclic coordinates reduces to the quadratures

J
t
8R 0
qm+s =- ~/3 dt + qm+s (s = 1, ... , n - m) . (7.16.7)
m+s
to
372 7. Lagrange's differential equations

Equations (4), (6) and (7) present the complete system of integrals of
the initial system of differential equations of motion with 2n arbitrary con-
stants. The presence of n - m cyclic coordinates allows one to reduce the
order of the system to 2m, the problem reducing to the integration of re-
duced system (5) together with n - m quadratures (7). It it necessary to
add that the number of cyclic coordinates depends upon the choice of gen-
eralised coordinates. For example, there are no cyclic coordinates when the
position of a particle in the central field is described by the Cartesian co-
ordinates x, y, z, whereas one cyclic coordinate (the longitude) exists if the
spherical coordinates are used, see the first example in Sec. 7.18.
Clearly, expressions for the cyclic coordinates can be written by means
of eq. (15.19) in the form

qm+s = J
to
(Nm+s - f M~+sqk)
k=1
dt + q!+s· (7.16.8)

In the case of the potential forces, subtracting the expression for the
potential energy II(q1' ... , qm) from both sides of eq. (1) and noticing that
8L
f3m+s = Pm+s = -8-·--' (7.16.9)
qm+s
we introduce the Routhian kinetic potential LR as follows

LR = R - II = [ L -
n-m
~ f3m+sizm+s .
1 (7.16.10)
s- qrn+r=frn+r

The differential equations of motion (5) take the form

C's
£." (L R ) = !!:.... 8L R _ 8L R = 0 (S = 1, ... ,m ) . (7.16.11)
dt 8qs 8qs
Provided that time t does not appear explicitly in the expression for LR
then repeating the derivation of Sec. 7.2, we obtain the following energy
integral for eq. (11)

(7.16.12)

It follows from eqs. (1) and (15.19) that function R can be split into three
terms
(7.16.13)
where R2 and R1 are homogeneous quadratic and linear forms in the gen-
eralised velocities q1, ... , qm, whilst Ro does not depend on them. Thus,
similar to (2.6) the expression for the energy integral is written in the form
R2 + II - Ro = h. (7.16.14)
7.17 Structure of the Routhian function 373

7.17 Structure of the Routhian function


Let us calculate the following sum
n-m

l: Pm+s'im+8
8=1
~
~
( -!:}-.--
8=1
aT:;* + -!:}-.--
uqm+s
au + -!:}-.--
aTr ) qm+8
uqm+s
.
uqm+8
2T';*+U+Ti*. (7.17.1)
Recalling definition (16.1) of the Routhian function and making use of eq.
(15.4) we obtain
R = T'; + U + T';* + T{ + T{* + To - 2T';* - U - T{*
or
R = T'; - T';* + T{ + To, (7.17.2)
where the cyclic generalised velocities should be replaced by their expres-
sions (15.17).
In the case of stationary constraints the latter expression takes the fol-
lowing transparent form
R=T'; -T';*, (7.17.3)
Le. the Routhian function is the difference between the quadratic forms
T2 in the generalised velocities corresponding to the positional and cyclic
coordinates. However when we replace the generalised cyclic velocities in
T:;* by their expressions (15.17), where B m +8 = 0 for the stationary con-
straints, the Routhian function in terms of the generalised velocities qk
(k = 1,2, ... , m) is no longer a quadratic form of the velocities but has the
structure prescribed by formula (16.13).
Let us perform this replacement in eq. (2), that is, for the general case
of non-stationary constraints. We have, cf. [66]

1"" A m+s,m+rqm+8qm+r
n-mn-m
T.**
2 = '2 ~ ~ ..
8=1 r=1

= ~ EE A m+8,m+r [Nm+8 N m+r - 2Nm+r fqkM~+8+


8=1 r=1 k=1

ff
k=ll=1
M~+sM;"+rqkqll = <1>0 + <1>1 + <1>2, (7.17.4)

Le. T:;* is the sum of three components, namely the function independent
of the generalised coordinates
1 n-mn-m
<1>0 = '2 l: l: A m+s,m+r N m+8 N m+r. (7.17.5)
8=1 r=1
374 7. Lagrange's differential equations

the linear form in these coordinates


m n-mn-m
<1>1 = - Lqk L L Am+s,m+rNm+rM!+s
k=1 s=1 r=1
(7.17.6)

and their quadratic form


1 m m n-mn-m
<1>2 = '2 LLqkqZ L L Am+s,m+rM!+sM;"+r' (7.17.7)
k=1 Z=1 s=1 r=1

Correspondingly, expression (2) for the Routhian function takes the form

(7.17.8)

and, in the case of stationary constraints,

(7.17.9)

We replace now N m +r in eq. (5) by their values due to relationships


(15.18). The latter yield
n-m
f3m+s - Bm+s = L Am+s,m+rNm+r' (7.17.10)
r=1

Using eq. (5) we obtain

aN0<1>0
m +s
= f3m+s - B m+s. (7.17.11)

Now referring to eqs. (4.2.9) and (4.2.12) we can write an expression for
the quadratic form <1>0 of Nm+s both in an associated form and in the linear
form

<1>0 ~ 'f''f'
s=1 r=1
Am+s,m+r (f3 m+s - Bm+s) (f3m+r - Bm+r)

1 n-m
'2 L (f3m+s - Bm+s) N m+s. (7.17.12)
s=1

It is worthwhile mentioning that as follows from (5) both the quadratic


form <1>0 and T2* have the same matrix A **, thus <1>0 has a positive definite
form.
The coefficients of the linear form <1>1 are transformed by eq. (10) yielding
instead of (6) the following
m n-m
<1>1 =- L qk L M!+s (f3m+s - Bm+s) . (7.17.13)
k=1 s=1
7.17 Structure of the Routhian function 375

Let us construct now the expression for the kinetic energy of the sys-
tem in terms of the generalised velocities corresponding to the positional
coordinates. To this aim we replace tim+s and R in the expression
n-m

T =R+L fim+stim+s (7.17.14)


s=1
by means of eq. (15.18) and (8), respectively, using eq. (13) and the bilinear
form of <Po given by the second line in (12).
We obtain
m n-m
T = T; - <P2 + T{ + Ltik L M~+s (fim+s - B m+8) + To-
k=1 s=1

~ ~ (fim+8 - Bm+s) N m+8 + ~ fim+s ( N m+8 - ~ tikM~+8) .


(7.17.15)

Replacing T on the left hand side by the sum

(7.17.16)

and Tl by the following expression


m n-m

Tl L Bktik + L B m+stim+8
k=1 8=1

T{+ ~ Bm+s (Nm+s- ~M~+Ak)' (7.17.17)

we arrive, after rearranging the terms on the right hand side of eq. (15), at
the following relationship

m n-m
= (T; - <P2) + T{ - L tik L M~+sBm+s + To
k=1 s=1
1 n-m
+ 2 L (fim+s + Bm+s) Nm+s.
8=1

By virtue of eqs. (12) and (8) this can be simplified to give

T2 = R2 + <Po· (7.17.18)
376 7. Lagrange's differential equations

It follows from this representation that

is a positive definite form of the generalised velocities (11, ... , rim correspond-
ing to the positional coordinates. One arrives at this conclusion from the
following considerations. Let ri2 denote the initial values of the generalised
velocities rik and let us assume that the cyclic momenta (3m+s equal the
initial values of the coefficients B m +8 , i.e.

at t=to Bm+8(q~, ... ,q~;to) = (3m+s (8=1, ... ,n-m).


(7.17.20)

Then, due to eqs. (15.18) and (15.17), the initial values of the cyclic gen-
eralised coordinates are
n-m
·0 _ ~ Mk 0 . t ) ·0
qm+s - - ~ m+s (0
q1"" ,qm' 0 qk' (7.17.21)
k=l
Under such initial conditions <1>0 = 0 and R2 = T 2 , that is R2 is positive.
Given that the initial moment as well as the initial values for the posi-
tional coordinates q~, ... ,q~ and the corresponding velocities ri~, . .. , ri~
are taken arbitrarily, proves the positive definiteness of the quadratic part
R2 of the Routhian function. To make this conclusion by analysing the
coefficients of the quadratic form in the form (19) is a much more difficult
task.
Let us summarise the aforesaid. The Routhian function is represented in
the form

R = T; - [T;*]q' -fm+s
'171.+8-
= R2 + R1 + Ra, (7.17.22)

where R2 being quadratic in ri1, ... , rim is positive definite, whilst the linear
part R1 due to eqs. (9) and (13) is equal to
m n-m

R1 = Lrik L M~+s(3m+8' (7.17.23)


k=l s=l

Here the coefficients M~+8 are determined in terms of the coefficients in


the expression for the kinetic energy T due to the rule (15.17). Finally, the
free term

Ra = -cI> =
o
-~2 ~~
~ ~
A m+s ,m+r(3m+s (3 m+r (7.17.24)
8=1 r=l
7.18 Examples 377

is a negative definite form of the generalised momenta. When U = 0 in


the expression for the kinetic energy T the cyclic momenta are, due to eq.
(15.12), the derivatives of T;* with respect to qrn+s. This means that the
quadratic form T;* is an associate form for - Ro
n-mn-m

T{* = ~ L L A rn+s,rn+r,8rn+s,8rn+r = -R~. (7.17.25)


s=1 r=1

The equations of motions for the positional coordinates are given by the
form
~ 8R
~
2 _ 8R
~
2 = Qk + fk + 8R
~
o (k = 1, ... ,r.n), (7.17.26)
dt uqk uqk uqk
where fk denote the gyroscopic forces (Sec. 7.3) corresponding to R 1 . R2
can be treated as the kinetic energy and - Ro as the potential energy. Ac-
cording to (25) the latter is equal to the kinetic energy of the hidden mo-
tions (when U = 0). In the case of prescribed potential forces and stationary
constraints the energy integral (16.14) says that the sum of the kinetic en-
ergy R2 and the corrected potential energy of the system II - Ro remains
unaltered. We encountered the appearance of gyroscopic forces while elim-
inating the cyclic coordinates in the third example of Sec. 7.9. In Hertz's
mechanics [37] the potential energy of the field of any force is treated as
the kinetic energy of the hidden motions.
Assume that we can observe only those motions which correspond to the
explicit coordinates. We establish i) a change in the values and distribution
of the system masses (as R2 does not coincide with T;), ii) appearance of
the gyroscopic forces, and iii) a change in the field of the potential force
due to the corrected potential energy. These effects describe the influence
of the hidden motions on the explicitly observable motions.
If Rl = 0 the gyroscopic forces are absent and the system is gyroscopically
uncoupled. By virtue of eqs. (23) and (15.18) this occurs if
Arn+s,k = 0 (s = 1, ... , n - r.n; k = 1, ... , r.n), (7.17.27)
i.e. when the expression for the kinetic energy T contains no products
of the cyclic velocities and the velocities corresponding to the positional
coordinates (for U = 0).

7.18 Examples
7.18.1 Motion of a particle in a central force field (Keplerian
motion)
The equation of motion in vectorial form is as follows

r.nr.. = --r f ()
r
r '
378 7. Lagrange's differential equations

where r denotes the position vector of the moving point referring to the
attraction centre 0, f (r) expresses the dependence of the force on the
distance. Its moment about point 0 is obviously equal to zero as
r
rx-f(r)=O.
r
Therefore

.. d . 0
r x mr = dt r x mr = ,

and the angular momentum of the particle about centre 0 is constant

KO = r x mi".

This means that the trajectory of the particle is a curve which lies in the
plane perpendicular to the vector KO and passes through the attraction
centre. We take the latter as the origin of the polar coordinate system r, 'P
in this plane. The expressions for the kinetic and potential energies take
the form

J
r

T = ~m (r2 + r2<p2), IT = f (r) dr. (7.18.1)


ro

The coordinate is cyclic and thus the corresponding integral

(7.18.2)

expresses the known law of areas implying the constant value of vector KO.
The Routhian kinetic potential is

LR = T; - T;* - IT = ~mr2 - !;2 + J


2

ro
r

f (r) dr, (7.18.3)

and the differential equations of motion takes the form

[r(LR)=mr-
mr
{l~3+f(r)=O. (7.18.4)

It admits the first integral


2 r
T; + T;* + IT = '!'mr 2 + {lip 2
2 2mr
+ J f (r) dr = h, (7.18.5)
ro

which is the integral of energy and can be obtained immediately from


(16.14).
7.18 Examples 379

If the particle position is determined using spherical coordinates with


spherical radius R centred at the attracting centre 0, the pole angle f) (the
complementary angle to the latitude) and the longitude ,x, then we can
decompose the vector v into components along the radius VR, along the
tangents to the meridian Va and the parallel v,\ such that

VR = R, Va = R{j, v,\ = R). sin f).

The expressions for the kinetic and potential energies are

J
R

T= ~m(R2+R2{j2 +R2).2sin 2 f)) , II(R) = f(R) dR. (7.18.6)


Ro

As ,x is the cyclic coordinate, then the corresponding first integral is given


by
aT 2· . 2
(3,\ = a'x = mR 'xsm f), (7.18.7)

and the Routhian kinetic potential is determined as

LR = 1
-m R (·2 + R2·2) f) - (3i - II. (7.18.8)
2 2mR2 sin 2 f)
The differential equations of motion are

mR - mR{j2 + II' (R) -


mR sm
f~ 2
f)
= 0, }
(7.18.9)
m (R2{j)· _ (3i cos f) = o.
mR2 sin3 f)
They admit the first integral

~m
2
(R2 + R2{j2) + II (R) + (3i 2
2mR2 sin f)
= h. (7.18.10)

This first integral could be written directly from eq. (16.14). By means of
this integral the quantities {j and f) can be removed from the first equation
in (9). We will return to the problem of Keplerian motion later.

7.18.2 Heavy top


We consider motion of a heavy rigid body having an immobile point 0
on axis Oz which is the axis of symmetry of the inertia ellipsoid at point
O. We introduce "half-moving" axes n, n', i~ (see Sec. 2.10 for detail) and
take the projections WI, W2, W3 of w on these axes as quasi-velocities. Then
setting

eO = A (nn + n'n') + Ci~i~, (7.18.11)


380 7. Lagrange's differential equations

A and C being equatorial and polar moments of inertia, respectively, we


obtain by means of eq. (4.7.4)

(7.18.12)

By virtue of eq. (2.10.4) we can write the kinetic energy in terms of Euler's
angles 'l/J, iJ, 'P in the form

(7.18.13)

The potential energy is

II = Mg(c = Mgz cos iJ, (7.18.14)

where (c denotes the coordinate of the centre of inertia along the upward
vertical O( and z stands for its coordinate along the axis 0 z of the body
symmetry. The cyclic coordinates are 'l/J and 'P and the corresponding inte-
grals are given by

A~ sin2 iJ + C (cp + ~ cos iJ) cos iJ = (3'1j;' C (if + ~ cos iJ) = (3'P.
(7.18.15)

The Routhian kinetic potential is then determined by the equality

(7.18.16)

and the integral of energy is

·2 ((3 - (3 cosiJ)2
AiJ + 'Ij; 'P + 2Mgz cos iJ = 2h. (7.18.17)
A sin 2 iJ
Putting cos iJ = u we come to the familiar differential equation for the
cosine of the nutation angle

(7.18.18)

We do not dwell on integration of this equation and analysis of the motion


as these problems are comprehensively considered in numerous courses on
mechanics and monographs, e.g. [30].
Of special interest is the case of a nearly vertical top (" sleeping top")
when angle iJ remains small. Euler's angles are not appropriate generalised
coordinates as angle 'l/J is not small and sin 2 iJ in the denominator of eq. (17)
leads to an additional difficulty. It is reasonable to take those two angles as
the generalised coordinates which remain small in the near vicinity of the
7.18 Examples 381

FIGURE 7.11.

vertical position of axis Oz. The ship angles introduced in Sec. 2.4 are suited
for this purpose. They are denoted here as a and f3 where a designates
the angle between axis Oz and plane O(~ and f3 the angle between the
projection of axis Oz on plane O(~ and axis 0(. The third generalised
coordinate is angle rp of the body rotation about axis Oz. The system of
three unit vectors n, n' , i~ "half-bounded" to the body is now defined as
follows: n lies in the plane O(~ perpendicular to the projection of axis
o z on this plane comprising angles f3 and 7r /2 + f3 with axes O~ and 0(,
respectively, n ' lies in the plane containing axes 0 z and Ory at the angle
a to the latter and i~ along the body axis, see Fig. 7.11. The projections
of the angular velocity w on these directions are expressed in terms of the
generalised velocities a, /3, <p as follows

Wn = WI = -a, Wn, = W2 = /3 cos a , Wi'3 =w3=<p+/3sina.


(7.18.19)

The table of the directions cosines for the half-bounded trihedron is given
by

ry (

}
~
n cosf3 0 - sinf3
(7.18.20)
n' - sin a sin f3 cos a - sin a cos f3
°1
13 cos a sin f3 sina cos a cos f3

The expressions for the kinetic and potential energies take the form

T = ~A (a 2 + l cos 2 a) + ~C (<p + /3 sin a f, n= Mgzcosa cosf3,


382 7. Lagrange's differential equations

indicating that only coordinate <p is cyclic. The integral corresponding to


the cyclic coordinate as well as the Routhian kinetic potential are as follows

f3<p =C (cp + ,8 sin a ) , }


1 .2 . 13 2
LR = 2A (a 2 + 13 cos 2 a) + f3<pf3 sina - 2~ - Mgzcosacosf3.
(7.18.21)

The energy integral has the form

A (a 2 +,82 cos 2 a) + 2Mgz cos a cos 13 = 2h. (7.18.22)

The first cyclic integral (15) obtained by means of Euler's angles states
that the projection of the resultant angular momentum on the vertical O(
is constant. The external forces acting on the top is the weight and the
reaction force of fixed point 0 and thus the moment of these forces about
the fixed axis O( is zero. When we adopt angles a and 13 as the generalised
coordinates we cannot find this integral using only expressions for T and II.
Taking into account that the projections of resultant angular momentum
on axes of the "half-bounded" trihedron n, n', i~

AWl = -Aa, AW2 = A,8cosa, CW3 =C (cp + ,8 sin a) = f3<p'


(7.18.23)

we obtain by projecting on axis O(

Aa sin 13 - A,8 cos a sin a cos 13 + 13 <p cos a cos 13 = 'Y, (7.18.24)

where 'Y is a constant value.


The equations of motion corresponding to the Routhian kinetic potential
(21) take the form

A (ii + iJ2 cos a sin a) - f3<piJcosa - Mgzsinacosf3 = 0, }

A (,acos 2 a - 2a,8sin a cos a ) + f3<pacosa - Mgz cos a sin 13 = O.


(7.18.25)

When a and 13 are small we arrive at the system of two differential equations

(7.18.26)

The solution of the system can be found, for example, in [56], see also
Subsection 7.9.3 of the present book. The integrals (22) and (24) are applied
to the analysis of the stability of the sleeping top in [20J.
7.18 Examples 383

7.18.3 System of two heavy tops


The first heavy top having a fixed point 0 1 rotates with angular velocity
w. Its inertia tensor at point 0 1 is denoted as 8 1 , the centre of inertia C 1
lies on the symmetry axis 01Zl with abscissa Zl. Point O 2 on axis 01Zl is
the point of support for the second heavy top rotating with angular velocity
!l. Its inertia tensor at point O 2 is denoted by 8 2 , the distance 0 1 0 2 is l
and the abscissa Z2 of the centre of inertia C 2 lies on the axis of symmetry
02 Z 2.
We use two "half-bounded" systems of axes introduced in the previous
example, axes n, n', i~ for the first heavy top and n*, n*', ij' for the second
one. The system has six degrees of freedom and the ship angles

are taken as the generalised coordinates.


In accordance with eq. (19) the quasi-velocities are the projections WI, W2,
W3 of vector wand the projections W4, W5, W6 of vector !l on axes of the
corresponding "half-bounded" system.
The kinetic energy of the first heavy top is given byeq. (12)

1
T1 = 2"A1
(2 2) 1 2
WI + W2 + 2"C1W3' (7.18.27)

The kinetic energy of the second heavy top is calculated by means of eq.
(4.7.7)

1 2
T2 = 2"M2V02 + M2V02' (0*')
!lX 13 Z2 + 2"!l.
1 8 2 ·!l. (7.18.28)

Here M2 denotes the mass, V0 2 the vector of velocity of the supporting


point and ij' Z2 the position vector of the centre of inertia of the second top
with the origin at point O 2 . Then we have

(7.18.29)

and moreover

Noticing that

we find
384 7. Lagrange's differential equations

and the kinetic energy of the system takes the form

T = ~ [A~ (w~ + w~) + A2 (W~ + W~) + ClW~ + C6W~] + (7.18.31)

B (WlW4n' . n*' + W2w5n· n* - W2W4n· n*' - Wlw5n'.n*) ,

where A~ = Al + M 212, B = M 21z2. In order to express T in terms of the


generalised velocities and generalised coordinates we use formulae of the
type (19) and the tables of direction cosines for the first and the second
tops. The result is

T = ~ [A~ (a~ + P~ cos2 a l ) + A2 ( a~ + P~ cos2 ( 2) +


C l (CPl +Plsinal)2 +C2 (<P2+P2sina2)2] +

B [al a2 (cos al cos a2 + sin al sin a2 cos {3) + PlP2 cos al cos a2 cos {3+

alP2 cos a2 sin al sin {3 - a2P l cos al sin a2 sin {3], (7.18.32)

where {3 = {32 - {3l. The potential energy is equal to

II Ql(C1 + Q2(C2
QlZl cos al cos {3l + Q2 (I cos al cos {3l + Z2 cos a2 cos {32) .

where (e 1 and (c2 are the abscissas of the centres of gravity of the tops
referring to the upward vertical 0 1 (. Next, Ql and Q2 denote the weight
of the first and second tops, respectively. Now

(7.18.33)

The cyclic coordinates are 'PI and 'P2 and the corresponding integrals are
given by

Cl W 3 = Cl (CPl + PI sinal) = ')'1, C2W 6 = C2 (CP2 + P 2 sin(2) = ')'2·


(7.18.34)

The Routhian function, due to eq. (17.22), takes the form

R = T2* - T** 1 [A'1 (.2


2 = 2" al + {3.21 cos 2 al ) + A 2 (.2
a2 + {3.22 cos 2 a2 )] +

B [al a2 (cos al cos a2 + sin al sin a2 cos {3) + P l P 2 cos al cos a2 cos {3+
alP2 cos a2 sin al sin {3 - a2Pl cos al sin a2 sin {3 + 'hPl sin al +
1 ·2
'Y2P2 sin a 2] - 2" b: - b:
1. 2
2" = R2 + Rl + Ro, (7.18.35)
7.18 Examples 385

and the integral of energy is as follows

R2 + II = h. (7.18.36)

We can construct another integral, namely the integral of the angular mo-
mentum about the vertical axis 0 1(. Calculation of the resultant angular
momentum KG, about the fixed supporting point is performed by means
of formula (4.8.10). For the first heavy top

(7.18.37)

In order to calculate Kf' it is necessary to have the expression for the


vector Q2 of the resultant momentum for the second top

Q2 = M 2v C2 = M2 (w x li~ +0 x zif). (7.18.38)

Taking into account eqs. (29) and (30) we obtain

Kf' = li~ X Q2 + z2i~' X MVG 2 + 82 .0


= M2l2 (wIn + W2n') + A2 (W4n* + W5n*') + C2W3i~' +
B [i~ x ( -W4n*' + W5n*) + i~' x (-WIn' + W2n)]. (7.18.39)

Now we need the projection K?' of the vectorial sum


KG, = K?' + Kf'
on the vertical. Denoting the unit vector of the upward vertical as k we
have
k -, =
X 13 n ,. . al cos f13 }
sm f3 1 - n sIn
(7.18.40)
k x ij' = n*' sin f32 - n* sin a2 cos f3 2

and then

k· (i~ x n*') (k -, ) . n '* =


X 13 sm (3 1 - n . n *' sm
n I . n *'. . al cos (3 1

cos al cos a2 sin (31 + sin al sin a2 sin (32

etc. We obtain then

K?' = - {A~ (WI sin (31 + W2 sinal cos (31) + A2 (W4 sin (32+
W5 sin a2 cos (32) + B [W4 (cos a1 cos a2 sin (31 + sin a1 sin a2 sin (32) +
Wj (cos al cos a2 sin f3 2 + sin al sin a2 sin (31) + W5 sin a1 cos (32 +
W2 sin a2 cos (31]} + C1W3 cos a1 cos (31 + C2W6 cos a2 cos f3 2 = "
(7.18.41 )
386 7. Lagrange's differential equations

where "I denotes a constant value. Under our choice of generalised coordi-
nates this integral is not a cyclic one. Using angles 191, 1/J1, <PI and 192, 1/J2, <P2
for describing the positions of the first and second tops, respectively, instead
of eq. (33) we would obtain

II = (Q1Z1 + Q2l) cos 191 + Q2Z2 cos 192,


where 191 and 192 denote the angles between the tops' axes and 0 1 (. The
expressions for the kinetic energy would contain angles 191 and 192 (similar
to a1 and (2) and the difference 1/J2 -1/J1 (similar to /3 1 and /32 in eq. (32)).
Thus, if we adopt
1
1/J2 = X + "2 u ,
then X becomes the cyclic coordinate as it does not appear explicitly in
the expressions for the kinetic and potential energies. The corresponding
integral ensures that the projection of the resultant angular momentum on
axis 0 1 ( remains unaltered. In this regard the ship angles undermine the
expression for the potential energy. However they considerably simplify the
analysis of the nearly vertical top.
In this case, retaining in the expression for R only terms of second order
of the assumed small quantities ai, /3i' ai, /3i we obtain

R = ~[A~(a~+/3~)+A2(a~+/3~)+2B(a1a2+/31/32)]+
"11/31 a1 + 'Y2/32a2, (7.18.42)

(7.18.43)

The equations of motion are

a1 + b1a2 - 81/31 - C1a1 = 0, /31 + b1/32 + 81 a 1 - c1/31 = 0, }


a2 + b2a1 - 82/32 - C2a2 = 0, /32 + b2/31 + 82a2 - C2/32 = 0,
(7.18.44)

where

The equations of motion can be reduced to a system of two second order


differential equations in the following unknown variables

(7.18.46)
7.19 Quasi-cyclic coordinates 387

which are complex numbers determining positions of the vectors i~ and ij'
in the plane parallel to Ol~TJ. These equations have the form

The particular solution is sought in the form

(7.18.48)

where Zl, Z2 and A are constant values. Substituting (48) into (47) yields
the system of linear homogeneous equations

Zl (A2 + SIA + Cl) + Z2blA2 = 0, }


(7.18.49)
Z2b2A2 + Z2 (A2 + S2A + C2) = 0,

having a nontrivial solution for Zl and Z2 provided that the determinant


is zero. This condition leads to the following equation fourth order in A

(7.18.50)

The particular solution exists for any of the four roots Ak, the values of Zl
and Z2 being related to each other by the condition
Zk1 A% + S2Ak + C2 (7.18.51)
Zk2 b2 A%

The general solution of the system of equations of motion (47) can be


written in the form
4
Zl = L Ckb1A%ei)\k t , (7.18.52)
k=l
The necessary condition under which the axes of the heavy tops con-
serve nearly vertical positions is that the absolute values of Zl and Z2 are
bounded. (The sufficient condition is more difficult and is beyond the scope
of the present book). This takes place only in the case when all of the roots
of eq. (50) are real-valued. Indeed, if a complex-valued root A exists, then
its complex conjugate 5. also exists. The real part of one of roots iA, i5.
turns out to be positive and the corresponding particular solution becomes
unbounded as time t progresses. The condition under which all roots of the
fourth order equation are real is very complex and is derived in [84].

7.19 Quasi-cyclic coordinates


A generalisation of concept of cyclic coordinates turns out to be useful
for some classes of dynamical problem. Let us assume that there exists
388 7. Lagrange's differential equations

generalised coordinates qm+ 1, ... ,qn which do not appear in the expression
for the kinetic energy and the generalised forces however the corresponding
generalised forces Qm+ 1, . .. ,Qn (in contrast to the cyclic coordinates) are
not equal to zero. These coordinates are referred to as the quasi-cyclic
coordinates. An example is a rigid body with a flywheel rotating about
an axis fixed in the body and having the centre of gravity on this axis. If
q1, ... ,q6 denote the generalised coordinates describing the body position
(e.g. the pole coordinates and Euler's angles) and <p is the angle of rotation
of the flywheel relative to the body, then this coordinate is quasi-cyclic if
a moment depending on the generalised coordinates qs is applied to the
flywheel (for instance, to control the flywheel rotation).
Let ql, ... ,qm be positional and qm+l,· .. ,qn quasi-cyclic coordinates.
The differential equations of motion are then split into two groups

(7.19.1)

Pm+s = Qm+s (q1, ... ,qm) (8 = 1, ... ,n - m) , (7.19.2)

with Pm+s denoting the quasi-cyclic momenta

Pm+s =
aT
-!l-'-- (8 = 1, ... ,n - m). (7.19.3)
uqm+s

They are not constant under the motion but all formal constructions of
Sees. 7.15-7.17 remain valid.
In this fashion we can resolve the system of linear equations (3) for quasi-
cyclic generalised velocities

(7.19.4)

construct the Routhian function

R = [T - L.:: Pm+sqm+s1
n-m
(7.19.5)
s=1 tim+r=i'ffl+r

by means of the rule (16.1), and write the differential equations of motion
in the form

(7.19.6)

!l aR = -qm+s (8 = 1, ... ,n - m). (7.19.7)


UPm+s .
7.19 Quasi-cyclic coordinates 389

The systems of differential equations (6) and (2) should be considered to-
gether. Determining the positional generalised coordinates, the correspond-
ing generalised velocities and quasi-momenta, we find the quasi-cyclic co-
ordinates from eq. (7) by means of n - m quadratures. The advantage of
Routh's method is the reduction of the problem to the system of differential
equations of (n + m) - th order and the further quadratures.
Calculation of the Routhian function is simpler than (5) and is done,
under stationary constraints, in the following way
R = T* - (T**)q' m.+s--fm.+s = R2 + R1 + Ro, (7.19.8)
where T* and T** denote the parts of the kinetic energy, T* depending on
the generalised velocities corresponding only to the positional coordinates
and T** depending only on the quasi-cyclic generalised velocities.
The quasi-cyclic momenta do not appear in the quadratic form R2 in
the generalised velocities (h, ... ,elm' This follows from expression (17.19)
of this form and rule (15.17) of construction of M!+s' The term
m n-m

R1 = LL M!+AkPm+s (7.19.9)
k=1 s=1
is a bilinear form in the generalised velocities and quasi-cyclic momenta.
Finally, Ro is the quadratic form of the quasi-cyclic momenta and is written,
due to eq. (17.24), in the form
n-mn-m

°-
R - -"21 '~" '~" Am+s,m+r Pm+sPm+r' (7.19.lO)
s=1 r=1
Thus, taking into account eq. (2) we have

n-m

rk - L M!+sQm+s, (7.19.11)
s=1
where r k denotes the generalised gyroscopic force. Equation (6) can be
written as follows

(7.19.12)
We consider now the case of the potential generalised forces correspond-
ing to the generalised positional coordinates

Qk = --
all (k = 1, ... ,m). (7.19.13)
oqk
390 7. Lagrange's differential equations

In this case equalities (5.3.12) yield the following

aQk = aQm+s = o. (7.19.14)


aqm+s aqk
Therefore, excluding the case of the constant generalised forces Qm+s the
system with quasi-cyclic coordinates must have non-potential forces.
Using the relationship

dRo oRo. oRo . oRo. oRo


dt = L +L L 7)qk + L -a-Qm+s,
m n-m m n-m
7)qk -a-Pm+s =
k=l qk s=l Pm+s k=l qk s=l Pm+s
(7.19.15)

we repeat the derivation of Sec. 7.2 with respect to the differential equations
(12) and arrive at the equality

(7.19.16)

As R2 does not depend upon the generalised momenta we have

a (R1 + Ro) oR .
(7.19.17)
- aPm+s = - 0Pm+s = qm+s,
and relationship (16) serves to determine the power of the non-potential
forces Qm+s in the system with quasi-cyclic coordinates
n-m d
N = L Qm+sqm+s = dt (R2 + II - Ro) . (7.19.18)
s=l
8
Other forms of differential equations of
motion

8.1 The Euler-Lagrange differential equations


These equations differing from Lagrange's equations by introducing quasi-
velocities instead of the generalised velocities, were derived by Boltzmann
[13] and Hamel [35] at approximately the same time. It was Hamel who
suggested the above name for these equations. The equations of motion
used by Voronets [91] also deal with the quasi-velocities, however their
form differs slightly from the Euler-Lagrange equations.
The Euler-Lagrange equations are mainly used for non-holonomic sys-
tems and they were suggested to this aim as is seen from the titles of [13]
and [91]. However their significance is not limited to these special problems
since these equations can considerably simplify the form and the process
of constructing equations of motion for holonomic systems, too. We will
have an opportunity to convince ourselves as to how fruitful it is to apply
the Euler-Lagrange equations to some problems of dynamics of multi-body
systems.
We proceed from the general central equation (6.4.17). All we need is to
expand its left-hand side

(8.1.1)

and replace 8T on the right hand side by the expression


392 8. Other forms of differential equations of motion

n (aT aT)
bT = ~ aws bw s + a7rs b7r s . (8.1.2)

Substitution into eq. (6.4.17) yields


n doT naT.
~b7rsdtaws + ~~(b7rs)
naT naT n naT.
~ ~bws + ~ 87r s b7r s + ~ Psb7r s + ~ ~ [(b7r s ) - bW s ] -

(8.1.3)

Cancelling out the underlined terms we arrive at the relationship

n [dOT
Lb7rs -d ~
n n
+ LL'Y~s~Wt + LE~~
aT n aT aT
- -;:;- - Ps = O.
1
s=1 t UWs ,,=1 t=1 UW" ,,=1 uW" u7rs
(8.1.4)
We assume that the redundant coordinates are not introduced. We will
consider the cases of holonomic and non-holonomic systems separately.
In the first case, all variations b7r s are independent and the consequence
of (4) is that all of the coefficients of these variations are equal to zero. We
obtain the Euler-Lagrange equations of motion
d aT n n aT n aT aT
dt -;:;- +L L 'Y~s -;:;-Wt + L E~ -;:;- - -;:;- = Ps (8 = 1, ... , n) ,
Ws ,,=1 t=1 W" ,,=1 W" 7r s
(8.1.5)
The kinematic relationships (1.5.25)
n+1
qs=LbskWk (8=1, ... ,n) (8.1.6)
k=1
expressing the generalised velocities in terms of the quasi-velocities need
to be added to the Euler-Lagrange equations. We obtain a system of 2n
first order ordinary differential equations in the same number of unknown
variables
(8.1. 7)
The equations simplify and take the form
d aT
dt ~
uWs
n n
+ L L 'Y~s-;:;--Wt
,,=1 t=1
aT
uW"
- aaT7r s = Ps (8 = 1, ... ,n), (8.1.8)
8.1 The Euler-Lagrange differential equations 393
n
tis = LbskWk (8 = 1, ... ,n), (8.1.9)
k=l
if the quasi-velocities are introduced by means of the homogeneous linear
form (1.5.1) with coefficients which do not depend on t explicitly and not
by means of more general relationships (1.5.22).
As mentioned in Sec. 1.5, in the case of non-holonomic constraints the lin-
ear forms of the generalised velocities are understood as the quasi-velocities
which are identically equal to zero by virtue of the equations for the non-
holonomic constraints. The latter have the form of eq. (1.5.6), or in the
more general case,
n
Ws = Lasktik + as,n+l = 0 (8 = 1, ... ,l) (8.1.10)
k=l
if we use eq. (1.5.22). Then we have
n
thrs=Lask8qk=O (8=1, ... ,l), (8.1.11)
k=l
with l being the number of non-holonomic constraints. The summations
over 8 and t in eq. (4) should be performed from l + 1 to n

As variations 87fs for 8 = l + 1, ... , n are independent, the expressions in the


brackets must equal zero. We obtain n - l equations of motion

(8.1.13)

whose number coincides with the numbers of degrees of freedom. The fol-
lowing n kinematic relationships
n+l
tis = L bskWk (8 = 1, ... , n) (8.1.14)
k=l+l
need to be added to these equations. We have altogether 2n - l first order
equations in the same number of the unknown variables

(8.1.15)
394 8. Other forms of differential equations of motion

In order to construct equations (13) it is sufficient to know only the


symbols 1'rs whose subscripts correspond to the number of quasi-velocities
which do not vanish due to equations for the non-holonomic constraints.
Equations (13) contain the derivatives of the kinetic energy with respect
to all quasi-velocities including those which equal zero due to eq. (10).
With this in view, the non-holonomic constraints are not considered in the
expression for T. They are taken into account only after taking derivatives
of T with respect to the quasi-velocities which are the quasi-momenta p;.
Recalling expression (4.1.19) for the kinetic energy, we can retain the terms
which are linear in the quasi-velocities

(8.1.16)

i.e. we do not write down the products and squares of these quantities
as they cancel out after calculation of the quasi-momenta. We notice also
that among the terms linear in quantities (16) there are products of these
quantities and Wl+l, ... ,W n . Clearly, these terms must be retained in the
expression for T.
Equations of motion (5) and (8) become Lagrange's equations (7.4.1) if
all quasi-velocities Ws are the generalised coordinates

Ws=qs (s=I, ... ,n), (8.1.17)

since, as pointed out in Sec. 1.9, all symbols 1'rs and c~ are zero whilst
the "derivatives of T with respect to the quasi-coordinates" become the
derivatives with respect to the generalised coordinates qs. However it would
be an error to think that the s - th Euler-Lagrange equation becomes the
s - th Lagrange's equation. To understand this, it is sufficient to note that
the number of equations (5) or (8) does not coincide with the superscript of
symbol 1'rs. But in this particular case of prescribing the quasi-velocities by
relationships (1.9.6) and (1.9.7) all symbols vanish if one of the symbols is
greater than m. The equations of motion break up into two sets, namely the
Euler-Lagrange equations for numbers 1, ... , m and Lagrange's equations for
m+l, ... ,n.
We mentioned above that, as a rule, the structure of the expression for
the kinetic energy in terms of the quasi-velocities is much simpler than that
in terms of the generalised velocities. This explains why the Euler-Lagrange
equations of motion are simpler in form and more symmetrical than La-
grange's equations for many classes of dynamical problem. The difficulties
due to calculation of the three-index symbols are not so considerable and
not principal in any case. In addition to this, this calculation must be per-
formed for chosen quasi-velocities once and for all.
The case of rotation of a rigid body about a fixed point is an apt illus-
tration of the above. The expression for the kinetic energy in terms of the
generalised velocities is given by expression (4.7.6). Applying Lagrange's
8.2 Examples 395

equations
d aT aT
dt a;P - a1j;
d aT aT
dt ail - a{)
d aT aT

if the right hand sides are given byeq. (5.2.9) we would obtain cumbersome
expressions with no symmetry. By using the expression for the kinetic en-
ergy T in the form (4.7.5), the three-index symbols (2.10.3) and the Euler-
Lagrange equations we arrive at the well-known Euler's equations for the
body rotation about a fixed point

81Wl + (8 3 - 8 2 ) W2W3 = ml, }


82~2 + (8 1 - 8 3) w3wl = m2, (8.1.18)
83W3 + (8 2 - 8 1 ) Wlw2 = m3,

expressing the theorem on change in the resultant angular momentum.


This explains the adopted name of the considered form of the equations of
motion. The symmetry and transparency of Euler's equations are combined
next with the Lagrange formal approach to the calculation process.

8.2 Examples
Using Euler-Lagrange equations we consider now the problems for which
we obtained the expressions for the three-index symbols in Secs. 2.10 and
1.10 and for the kinetic energy is Sec. 4.13.

8.2.1 Sphere rolling on a rough surface


Taking into account the above we can write the expression for the kinetic
energy (4.13.2) in the form

2
5 (-2
T = "2IM [7a WI
-2) + 5
+ w2 2 a 2-2 2 - -
W3 + aW2w4 -
2aWlw5
- - ] + ... , (8.2.1)

where the terms w~ and w~ are omitted as the corresponding terms in the
equations of motion vanish by virtue of the equations for the non-holonomic
constraints. It is also necessary to bear in mind that the quasi-velocities,
which do not vanish due to the equations for the non-holonomic constraints,
are numbered by the indices 1, ... , l and in eq. (1.13) by the indices l+ 1, ... , n.
To construct the equations of motion one needs the three-index symbols
with subscripts 1,2,3. They are given by formulae (2.10.11) and (2.10.20).
396 8. Other forms of differential equations of motion

Taking into account the equations of the non-holonomic constraints (2.10.19)


we can write

pi = 8T)
( 8- -_ ~M
5 a 2-
WI, (8.2.2)
WI W5=0
and

(8.2.3)

Recalling that WI, W2, W3 denote projections of the angular velocity vector
w on axes Oxyz fixed in space, we obtain expression for the elementary
work of the active forces applied to the sphere in the form
s;:'w
u = V . uro
s;:
+m 0
.
()
= TTs;:
quXO + V2UYO
s;:
T-T
+ m- 01 u7f1
s;:-
+ m2 u7f2 + m3 u7f3.
- 0 s;:- - 0 s;:-

However byeq. (2.10.19)


8xo = a8ir 2 , 8yo = -a8ir 1 .
The generalised forces corresponding to the quasi-coordinates are
P3 -0
=m3· (8.2.4)
The first Euler-Lagrange equation is given by
7 . 5 3 8T
[;Ma 2 WI + LL1'~18Wr = Pl.
r=lf=l
It is necessary to specify the non-zero indices with right subindex 1. They
are
4
1'31 = -a.
We obtain
7 2- 8T _ 8T _ 8T _ .
-Ma WI + -_-W3 - -_-W2 - -_-aW3 - PI = O.
5 8W2 OW3 8W4
By virtue of eqs. (2) and (3)
8T _ 8T _ 8T _
8W2 W3 - 8W3 W2 - 8W4 aW3 = M a
2 (7 2 ) __
[; - [; - 1 W2W3 = 0,

the Euler-Lagrange equations for quasi-velocities W2 and W3 are derived by


analogy. Finally we obtain the following three equations
7
[; M2
:"
a WI = m 1 - a 2,
-0 V
7
[; M2
:" -0
a W2= m 2
V
+ aI, (8.2.5)
2 .
-Ma
5
2 W3= fi/i .
8.2 Examples 397

The kinematic relationships (2.9.4) and equations for the non-holonomic


constraints (2.10.19) should be added to the obtained equations. Then we
have 8 first order differential equations for the same number of unknown
variables

Let us construct Lagrange equations with multipliers in the form (7.1.6).


The expression for the kinetic energy should by obtained in terms of the
generalised velocities. By virtue of eqs. (2.9.4) and (4.7.6) it has the form

1 2 12 2(·2 ·2 2· )
T = 2M (x6 + Yo) + 2"5 Ma {} + 7/J + ~ + 27/J~cos{} .

Equations for the non-holonomic constraints are

Xo - a (Bsin 7/J - ~sin {}cos7/J) = 0, }


(8.2.6)
Yo + a (B cos 7/J + ~ sin {} sin 7/J) = o.

In order to avoid mistakes in numbering coefficients aks we construct the


general equations of dynamics (6.3.4) which are as follows

M (iooxo + iiooyo) + ~Ma2 (,a + ~~sin {}) O{} +


~ M a2 (~ + <P cos {} - ~B sin {}) o7/J + ~ M a2 (<p + ~ cos {} - ~B sin {} ) otp -
- - -0 0 0
V1 oxo - V2oyo - m3 o7/J - mNo{} - m3 otp = O. (8.2.7)

Equations for the non-holonomic constraints yield

OXo - a (sin 7/Jo{) - sin {} cos 7/JOtp) 0,


oyo + a (cos 7/Jo{) + sin {} cos 7/JOtp) O.

Multiplying these equations by the constraint multipliers -AI and -A2,


adding the results to the left hand side of eq. (7), and equating the coeffi-
cients of variation of the generalised coordinates we obtain five equations
of dynamics

Mio = Al + VI, Miio = A2 + t:2,


~Ma2 (,a + ~~cos{}) = m~ - a (AI sin 7/J - A2 cos7/J) ,

~ M a2 (<p + ~ cos {} - ~B sin {}) = m? + a sin {} (AI cos 7/J + A2 sin 7/J) ,
"52 M a 2 (..7/J + <p cos {} - ~{}. )
sin {} = in?
(8.2.8)
398 8. Other forms of differential equations of motion

The first two equations yield Al and A2. Substituting them in the remaining
three equations we obtain

~Ma2 (.;9 + '¢0 cos '!9 ) + Ma (xo sin 'ljJ - Yo cos'IjJ)


= m~ + a (VI sin'IjJ - V2 cos 'IjJ) ,
~ M a2 (cp + ,;j; cos '!9 - ~0 sin '!9) + Yo sin 'IjJ)
- M a (xo cos 'IjJ (8.2.9)

= m~ - a sin '!9 (VI cos'IjJ + V2 sin 'IjJ) ,

~ M a 2 (;j, + cp cos '!9 - 00 sin '!9) = m~ .


The two equations (6) for the non-holonomic constraints should accom-
pany these equations and allow xo and Yo to be removed from eq. (9). The
resulting system of equations should be equivalent to system (5). The co-
incidence of the third equation in (5) and the third equation in (9) can be
proved immediately. It is more difficult to prove that the remaining two
equations (9) are a consequence of eq. (5). We do not intend to prove this
since the aim of the analysis was to show how difficult and obscure for the
further investigations Lagrange's equations with multipliers are even for a
simple problem with non-holonomic constraints.

8.2.2 Ring
By virtue of eq. (4.13.6), the expression for the kinetic energy without terms
w~ and w~ is given by

(8.2.10)

Here, due to eq. (2.10.14)

Wl=iJ, w2=~sin'!9, w3=0+'¢cos'!9,


and equations for the non-holonomic constraints (2.10.26) take the form

W~ = W4 + aW3 = 0, w; = W5 + aWl sin'!9 = 0, (8.2.11)

where W4 and W5 denote the projections of the velocity of the ring centre On
the nodal axis and the axis perpendicular to it in the fixed plane Oxy. Ac-
counting for the weight we obtain the following expression for the potential
energy

rr = Mgz = Mga sin '!9. (8.2.12)

Since in this case 7fl = '!9 we obtain


orr
PI = - 0'!9 = -Mga cos '!9, P2 = P3 = o.
8.2 Examples 399

We have

pi ( -aT)
- 3
= -Ma 2
WI,
*
P2 = -aT- = -1 M a 2W2,
aWl ws=O 2 OW2 2
pj = 2Ma2w3, P: = - Maw3, P5 = -Mawlsin'!?
The derivatives
aT aT aT
07["1 ' 07["2 ' 07["3

are zero due to the second equation of the non-holonomic constraint. What
remains is to write down the values of the three-index symbols whose right
subscript is equal to 1,2,3. We have, due to eq. (2.10.16), (2.10.28) and
(2.10.29)
'Y~l - cot'!?, 'Y~l = 1, 'Y~2 = cot'!?, 'Y~2 = -1, (8.2.13)
4 15 1 5 a 5 a
1'52 sin'!?' 1'42 = sin'!?' 1'32 = - sin'!?' 1'23 = sin'!?·
We obtain the system of differential equations

3.
2Wl- 12
2W2cot'!?+2w2W3 = -~cos'!?, g}
1.
2W2
+ 2W2Wl
1 cot-
Q
'U - WlW3 =
0
,
(8.2.14)

2W3 - WlW2 = o.
If instead of t we take'!? as the independent variables, then
Ws = W~Wl (8 = 1,2,3) ,
.a
as = WI and the prime denotes a derivative with respect to '!? The second
and third equations (14) take the form
W~ + W2cot'!? - 2W3 = 0, 2w~ = W2. (8.2.15)
Removing W2 we arrive at the equation
w~ + w~ cot'!? - W3 = o. (8.2.16)
Under the initial conditions
t = 0, (8.2.17)
the general solution of the differential equation (16) integrated by means
of the hypergeometric series has the form
W3 = F ('!?,'!?o,wg,wg) , (8.2.18)
from which W2 is obtained by differentiation. Now the first equation (14)
reduces to the form
3 d 2 9 1 2
'4 d,!?W l = -~ cos'!? + 2W2 cot'!? - 2W2W3,
yielding WI ('!?) as a quadrature. A further integration determines'!? (t), [1].
400 8. Other forms of differential equations of motion

8.2.3 Two-axle trolley


The kinetic energy is determined by expression (4.13.18) in which iJ and CPs
are given by eqs. (4.13.13)-(4.13.17). Due to (1.13) the equations of motion
are
d aT S S aT aT
dt -a +L L 'Y~7-a Ws - - a P7,
W7 r=l 8=7 Wr 7r7
daT ss aT aT
dt -a + LL'Y~s8w8 - -a Ps·
Ws r=l s=7 Wr 7rS

Recalling equalities (1.10.15) we obtain

~ aT _ aT Ws _ aT = P7 }
dt aW7 aWl a7r7 '
(8.2.19)
~ aT + aT W7 _ aT = Ps .
dt aws aWl a7rs
Now we need momenta pi, P7' Ps and then to take into account the equa-
tions for the non-holonomic constraints

(8.2.20)

We obtain

~ -aOT)
:~) 0 = P,lW7 sin X cos X + Ws// cos X,

= (p,+p,lsm .2
X ) W7+ W S// .
sm X, (8.2.21 )
W7 0

( ::) 0 = //lws + //W7 sin X

implying that the quasi-velocities W2, ... , W6 can be taken to be equal to


zero. With the help of eqs. (1.5.17) and (1.10.14) we find

aT = Ls br7 -aT = b47 -aT = 0,


-a )
a a
7r7 r=l qr X
s (8.2.22)
aT
-a
7rs
= Lb
r=l
aT
rs -
qr
a
aT
X
2.
= 8 = P,lw7smxcosX + //W7WSCOSX·

Here the expression (4.13.19) for T, calculated under the presence of all
constraints, was used. Inserting this expression into eq. (19) we arrive at
the equations of motion

(p, + P,l sin 2 X) 0.17 + P,lw7WS sin X cos X + //0.1s sinx = P7, }
(8.2.23)
//l0.1s + //0.17 sin X + //W7WS cos X = Ps·
8.2 Examples 401

Denoting Ws = X, W7 = V, where V is the absolute value of velocity of


the joint B linking the axles we rewrite these equations in the form

(J-L + J-Ll sin2 X) V + J-Ll Vxsinxcosx + vx sin X = P7, }

( X+ ~ Vsin X) = :I Ps· (8.2.24)

As follows from eq. (4.13.20), the value vI presents the moment of in-
ertia of the front axle reduced to the axis of joint B, PsOX denotes the
elementary work of the active forces in the virtual displacement, deter-
mined by variation of angle X, and the generalised force P s denotes the
steering torque. The active forces include the engine torque transmitted to
the drive wheels, frictional forces in the wheel axles, resisting force of the
air and force of rolling friction, P7 being its generalised force. The thrust
of the drive wheels due to the road grip are reaction forces. Its elementary
work is zero and it does not appear in eq. (24).
Let us consider such a motion in which P 7 = P s = 0 and the initial
conditions are given by

t = 0, V = Vo, X = Xo, X = Xo· (8.2.25)

The differential equations of motion admit two first integrals. One of them

. IV' .
X+ l sm X = Xo + lIV;'
0 sm Xo (8.2.26)

expresses the condition of constant angular velocity of the front axle whereas
the second states that the kinetic energy of the system (4.13.19) is constant

(J-L + J-Ll sin2 X) V2+ 2vXV sin X + IvX2


(J-L + J-Ll sin Xo) V0 + 2vXo Vo sinXo + IvX~'
2 2

This equality can be recast in the form

[J-L + (J-Ll - y) sin X] V2 + vI (X + ~ sin x)


2 2

[J-L + (J-Ll - y) sin2 xo] V02 + vI ( Xo + ~ sin xo) 2 (8.2.27)

Equations (26) and (27) yield the formula relating velocity V with angle X

V (
J-L+ J-Ll -
V) sm. 2 Xo
T
Va J-L + (J-Ll - y) sin 2 X .
(8.2.28)
402 8. Other forms of differential equations of motion

8.3 Rolling of a rigid body on a fixed surface


The necessary kinematic relationships are obtained in Sec. 2.16. Provided
that point 0 is taken as the centre of inertia the expression for the kinetic
energy in terms of the quasi-velocities can be written in the form

2T = Mv 2 + w· 8· w = Mix - w X pl2 + w· 8· w. (8.3.1)

Here X denotes the vector which reduces to zero due to the equations for
the non-holonomic constraints (2.16.11) and the angular velocity vector w
is expressed in terms of the quasi-velocity X4 and il" by eq. (2.16.5). We
obtain

2T = MX2 - 2MX· (w x p) + w· [8 + M (Ep· p - pp)]. w. (8.3.2)

The first term in this equality can be cancelled out. The second term should
be kept as the differential equations of motion contain derivatives with
respect to Xl' X2' X3 and these terms remain after setting Xl' X2' X3 equal to
zero. The third term is the homogeneous quadratic form in X4 and il' with
coefficients depending upon the Gaussian coordinates ql, q2. Therefore, we
take

2T = -2M X . (w x p) + 2T* , }
(8.3.3)
2T* =w· [8+M(Ep·p-pp)]·w.

To obtain the generalised forces we write down the expression for the ele-
mentary work

8'W = F . 8r + rn D . ()

and insert the expressions for the virtual displacement 8r of the pole 0
and () due to eqs. (2.16.14) and (2.16.18), 87r vanishing because of the
non-holonomic constraints. We obtain

8'W = (rn D - p x F) . (rn87r4 + la8qa).


The quantity rn D - p x F = M is the resultant moment of the prescribed
forces about the point of contact of the body and the plane. The generalised
forces are as follows

P4 = M . rn, P5 = M· h, P6 =M .}z, M = rn D - p x F. (8.3.4)

Expressions for the generalised velocities in terms of the quasi-velocities,


due to eqs. (2.16.13) and (2.16.9), take the form

ql = X5' q2 = X6' i9 = X4 + M (X5k*2 - X6k*1).


8.3 Rolling of a rigid body on a fixed surface 403

As T does not depend on iJ we have

aT* aT (aT) aT*


(aT)
aT
a7f5 = aql x=o aql ' a 7f 6 = aq2 x=o aq2·
(8.3.5)

We proceed now to constructing equations of motion in the form (1.13)

(8.3.6)

The summation over r is performed up to r = 4 as the three-index symbols


with superscripts 5 and 6 are zero. The equations are obtained under the
condition that Xl' X2, X3 are set to zero after the derivation.
Let us expand the double sums in equations (6) using the values of the
three-index symbols collected in the tables of Sec. 2.16

4 6 aT
LL1'~4ax Xt
r=lt=4 r

aT aT aT) 4 aT*
- ( -a k21 + -a k22 + -a k23 X4 + 1'56-a X5·
Xl X2 X3 X4

It follows from eq. (3) that

aT aT aT
-a k cd +a - ko.2 +a - ko.3 =-M{wxp)·ko.=M{wxko.).p.
Xl X2 X3

However, by virtue of eqs. (2.16.18) and (2.16.19),


404 8. Other forms of differential equations of motion

We obtain

-Mp· PailX4 + ~qaq,6 (b,61a2a - b,62a1a) ,

2 Me .,6
Mp· P1X4 - /i:::lX4Q (b,61 a21 - b,62 all)-
vial
aT* bl l b22 - bi2 .2
-
aX4 vTaI Q
'
4 6 aT
LL'Y~6&Xt M p. P2X~ - ~X4q,6 (b,61a22 - b,62a12) +
r=l t=4 Xr vial
aT* bll b22 - bi2 .1
-
aX4 vTaI Q
.
Here

e=m·p

is equal to the distance from the centre of inertia 0 to the plane on which
the body rolls. Additionally

where p denotes the absolute value of the radius vector p with the origin
at point 0 and the end at the contact point.
The final form of the equations of motion obtained by Voronets is as
follows

(8.3.8)

The asterisk in the expression for T is omitted, that is T denotes the kinetic
energy of rolling without slipping, i.e. expression (2) at X = o. Voronets
8.3 Rolling of a rigid body on a fixed surface 405

used equations of motion different from the Euler-Lagrange equations and


for this reason the present derivation differs from that by Voronets [91].
The general integral of the system (8) with five integration constants
yields the following functions of time

The right hand side of expression (2.16.19) becomes a known function of


time and angle iJ is given by a quadrature by introducing another integra-
tion constant iJo. Then we obtain the coordinates x, y of the trace L of the
contact point in the plane as time-functions by means of the quadratures of
eq. (2.16.10). These integrations add two other constants. The general solu-
tion will have eight constants altogether. Euler's angles, relating trihedron
Ox'y' z' fixed in the body to trihedron Oxyz fixed in space, are determined
in terms of ql, q2, '13. Indeed, the position of the "half-fixed" trihedron of
the unit vectors

(8.3.9)

with vectors k"" due to eq. (2.16.19), and the arc element da (defined by
equality da 2 = a"'f3dq"'dq(3) is then known at any time instant.
The table of directions cosines of the angles between trihedron Ox' y' z'
and the trihedron m, T, m x T is given by the matrix
dq'" dq'"
m1 P",l da k",l da
dq'" dq'"
/3= m2 P",2 da k",2 da
(8.3.10)
dq'" dq'"
m3 P",3 da k",3 da

whilst the rotation matrix making the trihedron Oxyz coincidental with
the half-fixed trihedron is given by
o 0 1
/31 = cos iJ sin iJ 0 (8.3.11)
- sin iJ cos iJ 0
Then the matrix

(8.3.12)
406 8. Other forms of differential equations of motion

presents the tables of directions cosines of the angles between axes OX' Y' z'
and Oxyz.
Next we construct the formulae for the coordinates xo, Yo, zo of the pole
o in the fixed coordinate system Oxyz. To this end, it is sufficient to project
the vectorial relationship

-=----+
on the above axes. Here r M = OM denotes the position vector of the
contact point and its projections on axes Oxyz are x, y, O. We arrive at the
formulae

Xo = x - (O!ux' + 0!2lY' + 0!3IZ'), }

Yo = Y - (0!12 X ' + 0!22Y' + 0!33 Z') , (8.3.13)


Zo = - (0!13X' + 0!23Y' + 0!33Z') = -m· P = -c.

The projections PI, P2 , P3 of the reaction force of the plane on axes


OX' Y' z' fixed in the body can be determined by means of the Euler-Lagrange
equations constructed for quasi-velocities XI, X2, X3. These equations are as
follows

d8T 63 8T
dt a + LL1'~saXt = Ps (8 = 1,2,3), (8.3.14)
Xs t=4 r=l Xr
where T is given in eq. (3). The summation over r is from 1 to 3 since
all three-index symbols with superscript 4 and one of the subscripts:::; 3
vanish. Using the tables of Sec. 2.16 we obtain for 8 = 1

d
dt (w x P)l + [m2 (w x P)3 - m3 (w x P)2] X4 +
PI
[h2 (w X P)3 - h3 (w x P)2] X5+[l22 (w x P)3 -l23 (w x ph] X6 = - M·
(8.3.15)

Here m s , hshs, (w x p)s denote the corresponding projections of vectors


m, ll,h, w x P on axes Ox'y' z'. The time-derivatives of quantities (w x p)s
are projections of vector (w x p)*, i.e. the derivatives of the vector w x P
with respect to axes fixed in the body. Thus, relationship (15) is the result
of projecting the vectorial equality

(w x p)* + (mn + l",q"') x (w x p) = - ~P


on axis Ox'. The first factor in the second term equals w, then

w* x P + w x p* +w x (w x p) = - M1 P. (8.3.16)
8.3 Rolling of a rigid body on a fixed surface 407

Clearly, the same result will be obtained if the theorem on momentum is


applied. Indeed, there is no slippage at the point where the body contacts
the plane, thus

Va +w x P =0, Va = -w x p,

where va is the velocity of the pole 0 coinciding with the centre of inertia.
Then

Mvo = M (~o +w x va) = P,

which coincides with eq. (16).


Vector w in eq. (16) should be replaced by its expression (2.16.18). The
angular velocity vector ~ is given by formula (2.16.24) and vector wx pis
determined as result of the following calculation

wxp* [mo + vk (b a1 P2 - ba2 Pl) qa 1 Pf3qf3


x

kf30qf3 - balqamql - ba2qamq2

or

(8.3.17)

The normal component of the reaction force is

or

(8.3.18)

whereby

(8.3.19)

We recall that m denotes the outward normal to the surface, i.e. N < 0 if
the body exerts pressure on the plane.
The frictional force of the rolling body

T = P - mN = T,p' (8.3.20)
408 8. Other forms of differential equations of motion

can be determined by the covariant components

T~, p.pI =M[(:';X PI ).p-rlilpI .k(3-W'PW'P+


I
OP ]
W2p oql

M [kl .po' + E C:a Ea,D - rlcl ~ (a(31a,2 -


vial
aj32a,d-

If:T (ba1 a , 2
ilY. - ba2 a , l ) (Eo' + 113 . pq,(3) op
+ W 2 p~, 1 (8.3.21)
vial uql
where Ea,D denotes the Levi-Civita symbols defined by (B.1.16). Rolling
without slipping occurs if the value of the frictional force does not exceed
the critical value f INI, i.e. under the following condition

(8.3.22)

8.4 The case of a body bounded by a surface of


revolution
It is adopted that axes Ox' y' z', 0 z' being the symmetry axis of the surface
of revolution, are the principal central axes of inertia. The principal central
moments of inertia are denoted as A, B, C. A may not equal B as the mass
distribution over the body may not be symmetric about axis Oz'.
The kinetic energy of the body is calculated using formula (3.3)

2T = W· [8 + M (Ep· p - pp)]. W =AIWi + BIW~ + CIW~ - M (w. p)2 ,


(8.4.1)
where
(8.4.2)
etc. whereas the scalar product w . p due to eq. (3.19) and the formulae of
Sec. 2.16 for the surface of revolution is given by
w . p = rlm . p + qa la . p = rlE + (p sin a (-z' sin a + v cos a) , (8.4.3)
where E = v sin a + z' cos a. The expression for the kinetic energy simplifies
if A = B and is equal to
2T = AIW 2 + (C - A - MZ'2) w~ - Mv 2 w; - 2Mz'vWeW3, (8.4.4)
where, by virtue of eq. (2.16.26),

w2 = 0,2 + (p2 sin 2 a + k 2s2, }


W3 = 0, cos a - (psin2 a, (8.4.5)
We = (0, + (p cos a) sin a.
8.4 The case of a body bounded by a surface of revolution 409

The potential energy of the weight is


II = Mgc = Mg (ll sin a + z' cos a) (8.4.6)
and the differential equations (3.8) of rolling of a heavy rigid body bounded
by a surface of revolution on the fixed plane take the form

! ~~ -M (v cos a - z' sin a) sO+


M (vsina + z' sin a) (-kv + sin a) srp = 0,
- - - - + M (vcosa - z"
doT aT )2
sma 0 -
dt as as aT (8.4.7)
M (vsina + Zl sin a) Orp sin a - aOkrpsina

°
= -Mgk (vcosa - Zl sin a) ,
d aT aT M ( . 'r\k
dt arp - a<p +
1
vsma + z cosa
)
VSH + aT· ·
aDs k sma = .

Here k denotes the meridional curvature. If A = B the coordinate r.p does


not appear 'mT an d term aT.
a<p III t hI "vallIS
e atter equatIon hes.
Instead of the second equation we can consider the energy integral
T+II=E. (8.4.8)
In the case A = B, by using independent variable s, we can cast the
remaining equations in the form

doT
- - - M( vcosa - z"sma
)r\~~+ }
ds aD
M (vsina + Zl sin a) (-kv + sin a) rp = 0,
! ~~ +
(8.4.9)
M (v sin a + Zl cos a) vOk + ~~ksina = 0.

These two first order differential equations determine quantities rp and


o as functions of s.Then the problem of determining S as function of time
from the energy integral is reduced to quadratures.
As an example, let us consider a body standing with the spherical end on
the horizontal plane, see Fig. 8.1. This case of integrability of the problem
of a body rolling on fixed plane was pointed out by Chaplygin in 1897 in
[17]. Chaplygin's equations differ from Voronets's equations used here.
We take A = B then
Zl = h + acosa, v = asina,
where h and a denote the distance from the centre of inertia 0 and to the
geometric centre of the spherical surface and its radius, respectively. Then
k = a-I, S = aa and equations (9) take the form
d aT M hO .
da aD + a sm a 0,

d aT M h
da or{; + a
(ah + cos a ) ~£sma
r\ . aT .
+ aD sma o. (8.4.10)
410 8. Other forms of differential equations of motion

FIGURE 8.1.

Removing M ahn from these equations we arrive at the relationship


d &T
da &(p
It admits the integral

&T
&(p -
(aIi + cos a ) &T
&n = const, (8.4.11)

which is a linear relationship between (p and n and enables one to express


these quantities in terms of each other. Eliminating now (p from the first
equation in (10) we arrive at the first order differential equation linear in
n which is always integrable in quadratures. The quantities (p and n will
be expressed in terms of a . Though the next step for obtaining the depen-
dence a (t) by means of the energy integral is reducible to quadratures, it
can be rather more fully analysed only under some specially chosen initial
conditions even when h = O. This special case, under the assumption the
sphere contains a rotating rotor was studied by Bobylev in [10], see also
[99] . It was shown by Chaplygin in 1903, [18], that the rolling problem of
a body bounded by the spherical surface at h = 0 and A =f:. B =f:. C can be
reduced to quadratures.
In the latter case
(8.4.12)
where

(8.4.13)
8.4 The case of a body bounded by a surface of revolution 411

and differential equations (10) yield

~~ = f3 = const, d~ :~ + Ma 2 0 sin 0: + f3 sin 0: = o. (8.4.14)

Eliminating cp from the second equation and rearranging the terms we cast
this equation in the form

80 A(C+Ma2 ) (f3) (8.4.15)


80: + AC + M a2 (A sin 2 0: + C cos 2 0:) 0 - A = o.
Integration yields

f3 cos 0:
rl
= - + "( ---;==;=====;;== (8.4.16)
VI + >. cos2 0:'
H
A
where "( is a new constant and>' is defined by the equality

C-A Ma 2
>.= C+Ma 2 A .
The first equation (14) is written now as follows

.
'P=--
1
sin 2 0:
(f3
-coso:+"(
A
coso:- A-C A)
VI + >. cos2 0:
. (8.4.17)

Let us consider a particular solution

a= 0, 0: = 0:0,

obtained by means of the relationship between the constants f3, ,,(, E, 0:0
derived from the energy integral. In this case the trace of the contact point
is a parallel circle 0: = 0:0. By means of eq. (2.16.25) we obtain

iJ = 0 + cp cos 0:0, iJ = (0 + cp cos 0:) t + iJo

and, due to eq. (2.6.10), the trace of the contact point is the circle

acp sin 0:0 . acp sin 0:0


x = Xo + O+ .'PcoSO:o sm iJ, Y = Yo - O .
+ 'PcosO:o
cos iJ.

Finally, 'P = cpt+'Po· The solution has seven constants 0:0, f3, ,,(, iJo, xo, Yo, 'Po,
the eighth ao = 0, i.e. the motion in question occurs if the vector of the
initial angular velocity lies in the meridional plane (w~ = 0).
In a less general case the trace of the contact point on the sphere is a
meridian. Then cp = 0 which is possible as follows from eq. (17) only if
f3 = "( = 0 and thus 0 = O. Since the geodesic curvature of the meridian
412 8. Other forms of differential equations of motion

is zero, it follows from eq. (2.16.25) that 19 = 0, i.e. {) = {)o and the trace
of the contact point on the plane is a straight line. The law governing the
change of angle a with time is obtained from the energy equation which
for cP = 0, n = 0 takes the form

2(E - Mga) .
A+Ma2 =ao,

thus, a = aot + ao· The solution has six constants ao, ao, <Po, {)o and two
constants xo, Yo defining the initial position of the contact point trace on
the surface. The motion is feasible if the vector of the initial angular velocity
is perpendicular to the meridional plane.
The particular solutions of the considered type (rolling over the parallel
circle, rolling over the meridian) are also determined with small variations
in the problem of rolling without slipping on the fixed surface of the body
bounded by an arbitrary surface of revolution (for A = B).
Let us return to analysing the motion of a rigid body bounded by a
spherical surface under the assumption that the centre of inertia of the
body does not coincide with the geometrical centre of the sphere, i.e. h =I=- O.
Instead of eq. (13) we obtain the following expressions for the coefficients
of the kinetic energy

An = C cos2 a + (A + Mh2) sin2 a, }


A12 = [(A - C) cos a + Mh (a + hcosa)] sin2 a,
A22 = [(Csin 2 a+Acos 2 a+M(a+hcosa)2)] sin2 a, (8.4.18)
A33 = A + M (a 2 + h 2 + 2ahcosa) .
Integral (11) is now given by

[Aasin2 a + C (h + a cos a) cos a] n +


[Aacosa - C (h + a cos a)] cpsin2 a = f31h, (8.4.19)

and the first equation (10) takes the form

d~ (Ann + A 12 CP) + Mahn sin a = o. (8.4.20)

The calculation simplifies if, instead of cP sin2 a, we enter the projection


W3 of the angular velocity vector w of the body on symmetry axis Oz' due
to eq. (5)

W3 = ncosa - cpsin2 a. (8.4.21)

Equations (19) and (20) are transformed to the form

Aan - [Aa'Y - C (h + a'Y)] W3 = f31h, (8.4.22)


8.4 The case of a body bounded by a surface of revolution 413

d
d'Y {[A + Mh (h+ a'Y)] 0 - [(A - C)-y + Mh (a + h'Y)] W3} = MahO,
(8.4.23)

where cos Q = 'Y.


Rewriting these equations in the form

dO dw 3
Aa- - [Aa'Y - C (h + a'Y)]- - (A - C) aw3 = 0,
d'Y d'Y
dO dw 3
[A + Mh(h+a'Y)]-d - [(A - C)-y +Mh(a+h'Y)]--
'Y d'Y
(A - C + mh2) W3 = 0

and removing ~~ we obtain the following first order linear equation

[AC + Ma 2A (1- 'Y2) + MC (h + a'Y)2] ~3 +


[CMa (h + a'Y) - AMa2'Y] W3 = O. (8.4.24)

We easily find that

(8.4.25)

Now we obtain from eq. (22)

(8.4.26)

The expression for the kinetic energy in terms of 0, W3 and it is transformed


to the form

2T = 1 ~ 'Y2 {[A + M (h + a'Y)2] 0 2 + (8.4.27)

1 ~2~2W~ - 2 [A'Y + M (h + a'Y) (a + h'Y)] OW3 + A33'l}.


Obtaining the energy integral and inserting expressions (25) and (26) for
the quasi-velocities 0 and W3 into it we reduce the problem to determining
'Y as a function of time to quadratures, the latter being difficult to interpret.
We restrict our attention to the analysis of motions for which the symmetry
axis deviates slightly from the vertical. Then we can take

(8.4.28)
414 8. Other forms of differential equations of motion

7] being small. Denoting the values of W3 and n for "y = 1 by ro we obtain

AM a 2 - M C (a + h) a )
(
W3=rO 1- C[A+M(a+h)2] 7]+ ... ,
(8.4.29)
n _ {_ [ [Aa - C (a + h)]2 M _ C - A] }
~ G - ro 1 [ ] 7] + .. . ,
AC A+M(a+h)2 A

with knowledge of the higher order terms not being needed.


Calculating the expression in braces in formula (27) up to the terms of
7]2 we have

Returning to angle a we obtain

[A +M (a + h)2] a 2a2 +

Cr2a 2
o
(1 _( 2)
12
+ r2 [C + M (a + h) a]2 a 4
0 A+M(a+h)2 4
or

2T= [A+M(a+h)2] a2 +r2 {[C+M(a+h)a]2 +C} a 2 +Cr2.


o A+M(a+h)2 4 0

(8.4.31)

The potential energy is equal to

II = Mg (a + hcosa) , (8.4.32)

so that the energy integral is written in the form

[A+M(a+h) 2] a2 + [r.J!.2 {C+M(a+h)a}2


2
r6
+C--Mgh] a 2 =const.
4 A+M(a+h) 4
(8.4.33)

Taking the time-derivative yields

a··2 + {r6
- [C+M(a+h)a]2 +-1 cr6 -4M9 h } a-
-0 . (8.4.34)
4 A+M(a+h)2 4A+M(a+h)2
8.5 Appell's differential equations 415

Provided that the centre of inertia is under the geometrical centre of the
sphere, i.e. h < 0, angle a varies harmonically with the frequency

k= { r~
4
[0+M(a+ h)a]
A+M(a+h)2
2
+ ~ Or~ -
4Mgh
4A+M(a+h)2
}1/2
(8.4.35)

The positiveness of the expression in braces is the necessary condition for


the stability of the vertical position for h > 0, that is when the centre of
inertia is over the sphere centre, see [41].

8.5 Appell's differential equations


We consider a system of particles whose configuration is described by
n independent generalised coordinates ql, ... , qn. Equations for the non-
holonomic constraints, if they exist, are written in the form
n
L akstis + ak = 0 (k = 1, ... ,l), (8.5.1)
s=l
where aks and ak are functions of the generalised coordinates and time.
Equations (1) should be resolved for 1 generalised velocities til, ... , til, that
is the deficiency of 1 x n matrix (7.10.5) should be zero. Then expressing
1 generalised velocities in terms of the remaining n - 1 ones we obtain the
system of equalities

tir = br,l+ltil+1 + ... + br,ntin + br (r = 1, ... ,l) . (8.5.2)

Equations relating variations of the generalised coordinates


n
L ak s8qs = 0 (k = 1, ... ,l) , (8.5.3)
s=l
allow the variations 8q1, ... , 8ql to be expressed in terms of n - l variations
8ql+1, ... , 8qn, the latter variations being independent. By analogy to (2) we
obtain
n-l
8qr =L br,l+s8ql+s (r = 1, ... ,l). (8.5.4)
s=l
Differentiating eq. (2) we find expressions for the generalised accelera-
tions
n-l n-l
iir =L br,l+siil+s +L br,l+stil+s + br . (8.5.5)
s=l s=l
416 8. Other forms of differential equations of motion

These equalities will be used to eliminate ii,. from expression (1.3.9) for the
acceleration vector of particle Mi. We have
a a
ari..qm + ... = "Lr=1 a
n I n-l
Wi = "L ari qr
.. + "L -a--ql+s
ri .. + ... , (8.5.6)
m=1 qm qr s=1 ql+s

where dots stand for the components which do not depend on the gener-
alised accelerations. Making use of relationships (5) we recast the latter
equality in the form

Denoting

(i=l, ... ,N;s=l, ... ,n-l) (8.5.7)

we obtain
n-l
Wi = L Ci,l+siil+s + ... (8.5.8)
s=1

and since the unwritten terms do not depend on the generalised accelera-
tions we arrive at the equalities
aWi
aiil+s = Ci,l+s (8.5.9)

which are used in what follows. It is easy to understand that the virtual dis-
placements 8ri are expressed in terms of the independent variables I5ql+s by
means of the linear relationships with the same coefficients Ci,l+s. Indeed,
using (4) we have

(8.5.10)
8.5 Appell's differential equations 417

Applying relationship (10) to rearrange the left hand side of the fundamen-
tal equations of dynamics (6.3.2), we obtain
N
Lmiwi' t5ri
i=1

(8.5.11)

Noticing that

N aWi a 1 N as
Lmiwi' ~ = ~2 Lmiwi 'Wi =~, (8.5.12)
i=1 ql+s ql+s i=1 ql+s
we arrive at the equality
N n-l as
L miwi . t5ri = L ~t5ql+S' (8.5.13)
i=1 s=1 ql+s
in which S, due to (4.10.1), is the energy of accelerations. As seen from
the derivation, the presence of the non-holonomic constraints is taken into
account in the expression for S since the generalised accelerations ib, ... , iiI
were removed from the expression for S with the help of these constraint
equations.
Now we need to transform the expression for the elementary work of the
prescribed forces, which is the right hand side of the fundamental equation
of dynamics,
N N n-l n-l
t5'W = L Fi . t5ri =L Fi . L Ci,l+st5ql+s = L QI+st5ql+s' (8.5.14)
i=1 i=1 s=1 s=1
Here
N
QI+s = L Ci,l+s . Fi (8 = 1, ... ,n -l) (8.5.15)
i=1

denote the generalised forces corresponding to the independent variations


of the generalised coordinates. According to eq. (7) they can be recast as
follows

or, due to eq. (5.1.3),


I

Ql+s = Ql+s +L br,l+sQr (8 = 1, ... ,n -l), (8.5.16)


r=1
418 8. Other forms of differential equations of motion

where Ql+s, Qr denote the generalised forces calculated under the assump-
tion that all variations are independent.
The fundamental equation of dynamics is now brought to the form

L ~ - Ql+s Dql+s = o.
n-l (8S
s=1 uql+s
_)
(8.5.17)

Since the variations Dql+s are independent we obtain Appell's equations of


motion
8S -
~ = Ql+s (8 = 1, ... ,n -l) (8.5.18)
uql+s

first suggested in 1899, [2]. Their number is equal to the number of de-
grees of freedom. They are second order differential equations for the gen-
eralised coordinates ql+l, ... , qn. However, in the general case they contain
all generalised coordinates and velocities. Along with the equations for the
non-holonomic constraints which can be written in either of the forms (2)
and (1) we have a system of n differential equations in the same number of
unknown variables. The order of the system is 2 (n -l) + l = 2n - l.
Let us recall that Lagrange's equations with multipliers in the case of l
non-holonomic constraints present a system of differential equations of the
order 2n + l.
Appell's equations are also applicable when no non-holonomic constraints
exist. It will be shown below that they fully coincide with Lagrange's equa-
tions ofthe second kind (7.1.4) for the holonomic systems. Of course, while
constructing the expression for S it is necessary to take into account only
the terms containing the generalised accelerations and there is no need to
encumber calculation with the terms without these accelerations.

8.6 Appell's equations in terms of quasi-velocities


Calculation of the energy of the accelerations and construction of Appell's
equations are considerably simplified when quasi-accelerations are used in-
stead of the generalised accelerations.
Let WI, ... , Wn denote quasi-velocities, equations for the non-holonomic
constraints being cast in the form

Ws=o (s=l, ... ,l). (8.6.1)

The velocity vector of a system particle due to (1.5.19) is given by

(8.6.2)
8.6 Appell's equations in terms of quasi-velocities 419

Taking the time-derivative we obtain the acceleration vector


n-l
"
Wi = ~ -a--Wl+s
ari .
+ ... , (8.6.3)
s=l 7f1+s

dots standing for terms independent of the generalised accelerations. The


consequence of eq. (3) is the following relationship

(8.6.4)

On the other hand, taking into account that, in the case of non-holonomic
constraints (1),

87f s =O (s=l, ... ,l), (8.6.5)


we obtain due to eq. (1.6.14)

n-l ar. n-l aw'


8ri = L -a t
s=l 7f1+s
87fI+s = L - a" ' 87fI+s'
s=l Wl+ s
(8.6.6)

Thus
N
"m·w··
~ ""
i=l
8r·" ="m·w··
~
N

" " , , -a'- "


i=l
n-l

~
s=l
aw'

Wl+ s
87f1+
s
="
n-l

~
s=l
87f1+
s
N
"mw··
aw.
--"
~ , " a'
i=l Wl+ s
n-l 1 N a
= "87f1+ - - - "m·w· ·w·
~ s aWl s 2~ "" t
s=l + "=1

or

(8.6.7)

S being the energy of the accelerations. By virtue of eq. (5.1.6) the ele-
mentary work of the prescribed forces can be represented by the equality
which is, due to eq. (5), written in the form
n-l
8'W = L P I+ s87fHs, (8.6.8)
s=l

where PHs denote the generalised forces corresponding to the quasi-coordi-


nates. They, as well as S, are calculated with account of the non-holonomic
constraints.
The fundamental equation of dynamics is now transformed to the form

as )
L
n-l (

s=l
- a ' - PHs
Wl+ s
87fHs = O. (8.6.9)
420 8. Other forms of differential equations of motion

By virtue of the independence of variations 87rI+s we obtain Appell's equa-


tions in terms of quasi-velocities

8S
~ = Pz+s (s = 1, ... ,n -l). (8.6.10)
UWI+ s

This system of n-l first order equations contain the first time-derivatives
of quasi-velocities Wl+ s , the quasi-velocities and all generalised coordinates
ql, ... , qn and need to be completed by n first order equations (1.5.25)

n-l+l
qr = L br,I+kWI+k (wn+l = 1), (r = 1, ... ,n), (8.6.11)
k=l

determining the first time-derivatives of the generalised coordinates in


terms of the quasi-velocities.
Clearly, Appell's equations (10) are applicable to constructing the equa-
tions of motion for holonomic systems.
As will be shown below, Appell's equations (10) in no way differ from the
Euler-Lagrange equations (1.13). Application of either of these approaches
is simply a question of convenience of calculation. Use of the Euler-Lagrange
equations presumes the determination of three-index symbols in advance.
The kinetic energy should be found without account of non-holonomic con-
straints which complicates the structure of the expression and the very
construction of equations requires special attention to indices. While deal-
ing with Appell's equations the main difficulty is to calculate the energy
of the accelerations. One should be attentive in order not to overlook the
terms with quasi-accelerations. While considering the non-holonomic sys-
tems, the matter is facilitated by the opportunity to take into account
the presence of these constraints. The significance of the rules (4.10.4) and
(4.10.12) for construction of the energy of the accelerations S with the help
of the kinetic energy T should not be overestimated because the applica-
tion of the second rule requires knowledge of the three-index symbols and
an expression for T obtained under removed constraints, while the appli-
cation of the first one for construction of Appell's equations in the form
(5.18) reproduces the derivation that should be performed for Lagrange's
equation ofthe second kind (with non-holonomic constraints being absent).
The rules for construction of S, suggested in Sec. 4.11, have considerable
importance for the problems of rigid body dynamics. Appell's equations
are easy to memorize and write down once S has been constructed.
The systems subject to nonlinear non-holonomic constraints are not con-
sidered in the present book. Appell's equations [2] as well as the modified
Euler-Lagrange equations [36] are applicable to such systems.
8.7 Explicit form of Appell's equations. Chaplygin's equations 421

8.7 Explicit form of Appell's equations.


Chaplygin's equations
We start from representation (4.10.6) for the kinetic energy of accelera-
tions. With non-holonomic constraints (5.1) being absent, Appell's equa-
tions (5.18) according to (5.1) take the form

as
aqr = Qr (r = 1, ... ,n). (8.7.1)

Using eq. (4.10.6) we obtain


n n n
as
aqr L Asrqs + L L [s, k; r] qsqk +
s=1 s=1 k=1

However, recalling eqs. (7.4.1) and (7.3.5) it is easy to notice that the
right hand side of eq. (7.3.5) is nothing else than an expanded form of the
Eulerian operator £r (T). Equations (1) can be also represented in the form

as
-a"
qr
= £s (T) = Qr (r = 1, ... , n) . (8.7.3)

Thus, in the case of holonomic constraints, Appell's equations are identical


to Lagrange's equation of the second kind. Nothing other could be expected
as their right hand sides represent the same generalised forces.
Given constraints (5.2), we obtain by differentiation

qr = br,I+1ql+1 + ... + br,nqn + . . . (r = 1, .. , , n) , (8.7.4)

where dots designate the terms which do not contain the generalised accel-
erations. Energy of the accelerations is now dependent upon the generalised
accelerations qlH, appearing in eq. (4.10.6) explicitly as well as appearing
in S in terms of qr' For this reason

as (as) I as aqr (as) I as


aqlH = aql+k + ~ aqr aqlH = aqlH + ~ br,l+k aqr
or

(8.7.5)
422 8. Other forms of differential equations of motion

and by using eq. (5.16) Appell's equations (5.18) can be cast in the form
I
£l+k (T) + L br,l+k£r (T)
r=l
I
Ql+k +L br,I+kQr (k = 1, ... ,l) . (8.7.6)
r=l
It is easy to prove that this form of Appell's equations is identical to the
equations of motion (7.10.9) obtained by removing the constraint multi-
pliers from Lagrange's equations. It suffices to prove that the following
equalities

br,l+k = -NT+k (r = 1, ... ,lj k = 1, ... ,n -l) (8.7.7)

hold true. Indeed, the expressions in (5.2) are obtained from the system of
equations
I n-l
L aqrqr = - L aq,l+kql+k - aq (q = 1, . " ,l) (8.7.8)
r=l k=l
with the following matrix of coefficients

and the determinant la*1 = la~l. The algebraic adjunct a qr of element aqr
of the q - th row and the r - th column of this determinant is equal to
the algebraic adjunct a qr of element a qr of the q - th row and the r - th
column of the determinant la~ I of the transpose matrix a~. The solution of
the system of equations (8) has the form

n-l I 1 I qr
._~. '" qr '" a (8.7.9)
qr - - ~ql+k6Ia*la aq,l+k - 6aqla*I'
k=l q=l q=l

thus

(8.7.10)

and it is sufficient to refer to formula (7.10.10) in order to prove equality


(7).
8.7 Explicit form of Appell's equations. Chaplygin's equations 423

Instead of the described procedure for constructing Appell's equations


one could first eliminate all Qr (r = 1, ... ,l) from expression (4.10.6) for
the energy of the accelerations by means of eq. (4) and then differentiate
with respect to the remaining generalised accelerations Ql+k. This is implied
in the form (5.18) of Appell's equations which is written now as follows

~
as -
= Ql+k (k = 1, ... ,l), (8.7.11)
uql+k

S denoting the expression for the energy of accelerations after removal of


quantities Qr.
Let T denote the result of eliminating the generalised velocities qT) given
byeq. (5.2), from the expression for the kinetic energy. However it would
be a grave error to assume that under non-holonomic constraints

as
~=£l+k T (-) .
uql+k

If it were the case, then in accordance with eq. (6), all the differences

(8.7.12)

would be zero. Let us prove that they differ from zero. Following Chaplygin
[17] we perform this calculation under the assumption that all constraints
are stationary and the generalised coordinates ql, . .. ,ql do not appear in
the expressions for T and coefficients br,l+k' Then we have

and

aT aT
L -;:;-:-br,l+k,
I
~+
uql+k r=l uqr
d aT d aT I doT I aT.
dt ~
uql+k
+L
r=l
br,l+k dt ~ + L ~br,l+k'
qr r=l qr
I n-l
aT '"' aT ,"" abr l+m
--+~-~ql+m ' .
aql+k r=l aqr m=l aql+k

Now accounting for that

"s (T)
C' = .!!:... aT
dt aqr (r = 1,..., l) .
424 8. Other forms of differential equations of motion

we obtain
I
CI+k (T) +L br,I+kCr (T) +
r=l

aT (. n - l
L~
r=l
I

qr
br,l+k - L
m=l
Obr,l+m .
a
ql+k
)
ql+m· (8.7.13)

But
n-l
. ~ obrl+m .
br,l+k = ~ a' ql+m,
m=l ql+k
so that due to eqs. (6) and (5.15) we obtain

_
-
Q- l+k +~ aT ~ (Obr,l+k
~-. ~ --- - Obr,l+m).
ql+m
r=l oqr m=l oql+m Oql+k
(k = 1, ... ,n -l) (8.7.14)
This system of differential equations of the system subject to non-holonomic
constraints is identical to Chaplygin's equations. As one can see, in the case
of Lagrange's equations eliminating the generalised velocities qr by means
of constraint equations (5.2) from the kinetic energy T a priori would lead
to an erroneous result. If Lagrange's equations are used in such a way, they
should be completed by following correction terms

(k = 1, ... ,n -l) .

(8.7.15)
These terms vanish when
Obr,l+k _ obr,l+m = 0
oql+m Oql+k
(
r = 1, ... ,;1 m, k = 1, ... ,n -
l)
.

But in this case there exist functions fr (ql+l, ... qn) such that

br,l+k= !:l0fr (r=l, ... ,l;k=l, ... ,n-l)


uql+k
and the constraint equations (5.2) will be integrable (as the constraints
were assumed to be stationary) since
. ofr. ofr . f·
qr = -!:l--ql+l
uql+k
+ ... + ~qn
uqn
= r

or
qr = fr (ql+l, ... qn) (r = 1, ... ,l) .
The coordinates ql, ... ql are redundant. But in this case a priori removal
of qr from expression for T is justified and these coordinates could have
been omitted from the very beginning.
8.8 Applications to non-holonomic systems 425

8.8 Applications to non-holonomic systems


Let us return to the three examples considered in Sec. 8.2 and prove that
the same equations can be derived by Appell's approach.

8.8.1 Sphere
The energy of accelerations of a sphere rolling on fixed plane without slip-
ping is, due to eq. (4.13.3), given by

where we have omitted the terms corresponding to the equations of the


non-holonomic constraints, i.e.

The left hand sides of Appell's equations (6.10) are equal to

8S 7 2 ~ 8S 7 2 ~
-- = -Ma W2 -- = -Ma W3. (8.8.1)
. 5 ' "" ~ 5
8W2 UW3

The generalised forces are determined by expressions (2.4). Finally we ar-


rive at equations of motion (2.5).

8.8.2 Ring
The energy of accelerations is given by formula (4.13.12) which yields

(8.8.2)

The potential energy is due to eq. (2.12). Using this equation we can obtain
expressions for the generalised forces and then write down the equations of
motion (2.14).

8.8.3 Two-axle trolley


The left hand sides of the differential equations of motion (2.23) are ob-
tained by differentiating expression (4.13.23) for the energy of acceleration
with respect to W7 and W8.
426 8. Other forms of differential equations of motion

8.8.4 Chaplygin's equations for the problem of a rolling sphere


As shown in Sec. 8.2 the expression for the kinetic energy T in terms of the
generalised velocities and the equations for the non-holonomic constraints
have the form

1 12
T = 2M (xo2 + Yo)
2
+ 25 '19 + 't/J + cp + 2cp't/J cos '19 ,
Ma 2(·2 ·2 2 . )
(8.8.3)

Xo a (~sin't/J - cp sin '19 cos 't/J ),


Yo -a ( ~ cos 't/J + cp sin '19 sin 't/J) . (8.8.4)

Chaplygin's equations are applicable since the constraints are stationary


and xo, Yo appear neither in the expression for T nor in the constraint
equations.
Removing Xo and Yo from T by means of the constraint equations we
obtain

- 1 [7
T = 2M a2 5'19·2 + cp2 sin2 '19 + 5 (.2 2
't/J + cp + 2cp't/J cos '19 .
2 . )]

Adopting the following numbering of the generalised coordinates

we obtain

The generalised forces 02+r given byeq. (5.16) are equal to

03 m3, 04 = mN + a (VI sin 't/J - V2 cos 't/J) ,


05 m3 - a (VI cos 't/J + V2 sin 't/J) .
Now we need the values
8.8 Applications to non-holonomic systems 427

Their expanded expressions are collected in the following table

II II r = 1 r=2 r=3 II
-a (11 cos 'lj;+ a ( ~ sin 19 sin 'lj;
8=1 a cos 'lj; ( ~ + cp cos 19 )
cp sin 19 sin 'lj;) -11 cos iJ cos 'lj; )

-a (11 sin 'lj;- -a ( ~ sin 19 cos 'lj;


8=2 a sin 'lj; ( ~ + cp cos 19 )
cp sin iJ cos 'lj; ) +11 cos iJ sin 'lj; )

Since
aT . aT
a('iI = Mxo, a¢.2 = Myo,

we obtain the following values of the double summations on the right hand
sides of equations (7.14)

r = 1 °
r=2 -Ma2cp(~+cpcos19)sin19,
r=3 Ma211(~+cpcos19)sin19.
After simplification Chaplygin's equations take the form

2
SMa 2 (..'lj; + .)
(jJcos19 - cp19cos19 = m3,

~ Ma 2 (~ + cp~ sin 19) = mN - a (VI sin'lj; - V2 cos 'lj; ) ,


(8.8.5)
M a2 (..
tpsm . 2 00 2 .. 2 ~i. . 00 00 7. oQ • 00)
u+Stp+S,!-,smucosu-Stpusmu

= m3 - a (VI cos'lj; + V2 sin 'lj; ) .

Clearly, equations (2.9) take the same form when:To and Yo are replaced by
their expressions obtained by differentiating equalities (4). Equations (5)
are reduced to the simple and compact form (2.5) by projecting onto axes
Oxyz with the aid of formulae (5.2.15).
Using formulae (7.10.15) we construct expressions for the generalised
constraint forces

Qi = M:To - VI, Q2 = Myo - V2 , Q;; = 0, }


Q4 = a [- M (:To sin 'lj; - Yo cos 'lj;) + ~ VI sin'lj; -_ V2 cos 'lj; )] , (8.8.6)
Q'5 = a [M (:To cos 'lj; + Yo sin 'lj;) - (VI cos'lj; + V2sin 'lj; )] .
The complexity of the equations of motion (8.5) and calculation for such
a simple problem as the rolling of a sphere explains why Chaplygin did
428 8. Other forms of differential equations of motion

not use eq. (7.14) while solving particular problems. He derived it with the
aim of preventing an error which could occur when removing" excessive"
generalised velocities from the expression for T and indicated the form of
the correcting terms (7.15).

8.8.5 Plane motion of a particle


Let us consider the following instructive and simple example for applying
the Euler-Lagrange and Appell's equations. Let us agree, as suggested by
Appell, to describe the plane motion of particle M by means of its position
vector r = oM and the double area a of the sector OMoM swept over by
the position vector measured from the initial position OM~. Denoting angle
MoAM by <p we obtain the expression for the double sectorial velocity
(8.8.7)
so that the kinetic energy of a particle of unit mass is

1 (.2 + r<p
T =-r r +-
1 (.2
2.2) = - 0- 2 ). (8.8.8)
2 2 r2
Relationship (7) is a non-integrable equation defining the quasi-velocity
0- in terms of the generalised velocity cp. Designating
WI = r, W2 = 0-, 87r1 = 8r, 87r2 = 8a = r 28<p,
we have

(87r2t - 8W2 = 2rr8<p - 2rcp8r = ~r (W187r2 - W287r1)

and, byeq. (1.8.4), we find


2 2 2
'1'12 = -'1'21 = -. (8.8.9)
r
The elementary work of the force applied to the particle is
Fcp
8 W = Fr 8r + Fcpr8<p = Fr87r1 + -87r2.
I

r
Now we construct the equations of motions by means of (1.13)
d aT 2 oT oT
-dt --
aWl
+ 'Y21--W2
OW2
- --
07r1
d {)T 2 {)T
dt OW2 + '1'12 OW2 WI r
Calculation yields

.. 20- 2 0- 2 ~ if 2ro- 20-r 1


r- 7 + r3 = n r2 - 7 + 7 = -:;: Fcp.
8.8 Applications to non-holonomic systems 429

Hence we obtain the following equations of motion


·2
•• (J F (8.8.10)
r- 3 = r, a=rF'P.
r
Construction of Lagrange's equations by means of the kinetic energy (8)
would lead to incorrect equations since the constraint equation (7) is not
integrable.
Noticing now that
·2
Wr =
.. .2
r - rep =
..
r - 3'
(J
w'P = -1(2.)- a
r 'P =-,
r r r
we obtain expression for the energy of accelerations

8 -- -
1(
2
r··2 - 2··
r(J-
.2 + . 2)
r3
-(J
r2

and Appell's equations (6.10) yield the above equalities (10).

8.8.6 Friction gear


This example of a system subject to non-holonomic constraint is given
in [72]. Let us construct Appell's equations of motion for the mechanism
shown in Fig. 8.2. The mechanism transmits rotation of shaft 1 to drum
C fitted rigidly on shaft 2. Disc A is rigidly mounted on shaft 1, whilst
wheel B can rotate freely about shaft 3 but cannot move along the axis of
this shaft. The centrifugal governor P rotates together with drum C on the
same shaft, and displacement of its clutch D is transmitted to shaft 3 by
means of a cable. Under change in the angular velocity <P2 of shaft 2 (due
to acceleration or deceleration) the wheel B moves together with shaft 3
on disc A (to its centre or in the opposite direction). It is supposed that
the initial value of <P2 is restored as a result. The differential equations of
motion of the system are required.
The position of the system can be defined by three generalised coordi-
nates, namely the rotation angles 'PI' 'P2 of shafts 1 and 2 and the distance
x of the coupling D from joint L. The distance p = 8Q is expressed in
terms of x. This relation can be obtained if we notice that the sum of DID
and 02K is constant. Then we have

and, as OIL, 0 2 8, KQ are constant, the difference

x-p=c (8.8.11)

is constant, too.
430 8. Other forms of differential equations of motion

FIGURE 8.2.

As the circumferential velocities of the contact points of wheel B with


disc A and drum C are equal (no slipping is assumed), we obtain the
equation for non-holonomic constraint

(8.8.12)

The system has two degrees of freedom.


We proceed to calculate the energy of the accelerations retaining only
terms with the generalised accelerations. The energy of accelerations of disc
A, drum C, wheel B and coupling D are, respectively, equal to

(8.8.13)

Calculation of the kinetic energy of the spheres is more difficult. Denoting


the angle between rod LN and the axis of shaft 2 as a we have

x = 2l cos a.
Let us consider the projections of acceleration W N of the sphere centre on
three orthogonal directions which are the directions of the rod LN, perpen-
dicular to it in the plane of the centrifugal governor, and perpendicular to
this plane. The first projection does not contain generalised accelerations,
the second, as follows from Fig. 8.2,

lii - l<p~ sin a cos a ,

is the Coriolis acceleration and the third one is


8.8 Applications to non-holonomic systems 431

Thus, the energy of accelerations for both spheres has the components

~2mNl2 (&2 - 2&<p~ sin a cos a + cP~ sin2a + 4CP2<P2a sin a cos a) .
(8.8.14)
The generalised acceleration CP1 is removed from the expression for the
energy of accelerations by means of the equation for the non-holonomic
constraint (12), to give

R ( -<P2)· -_ -R (..'P2 - -
P<P2
-) ,
P P P

~2 (cp~ _2~CP2<P2) + ...


The justification of this procedure was discussed in Sec. 8.7. We have

x ..
a=-
(x i: cos a )
2
+----."....-
-2lsina' 2lsina 4z2 sin a
3 ,
x2 xi: 2 cos a

4z2 sin2 a + 4z3 sin4 a + ...


Now adding eqs. (13) and (14) we obtain S in the form

S =
1 .. 2[R2
-2'P28A 1
2+8c+8B2+-mN (2 R2
4l-x 2)]
(x - c) a 2

(8.8.15)

We determine now the generalised forces. Let <pg and Xo denote respec-
tively the values of <P2 and x under a stationary rotation. In this case the
spring of the governor has length h + 8 1 greater than the natural one,
whereas the spring of wheel B has length l2 - 82 less than the natural one.
Under a deviation from the stationary rotation the lengths of the springs
become
h + 81 + X - Xo, l2 - 82 +p- Po = l2 - 82 +X- Xo·

The changes in length, referring to the natural states, are 8 1 + x - Xo and


82 - x + xo, respectively, so the potential energy of the elastic forces up to
a constant term is

II
432 8. Other forms of differential equations of motion

The elementary work of the moments ml and m2 applied to shafts 1 and


2, respectively, is due to (12) given by

b'W = mlb'Pl + m2b'P2 = (ml ~


x-c
+ m2) b'P2

and the generalised forces corresponding to the coordinates 'PI and 'P2 are
R
ml--+m2,
x-c
- (CI + C2) (x - xo) - (clbl - C2b2).
The differential equations of motion obtained by means of scheme (5.18)
have the form

(8.8.17)

(8.8.18)

In order to prove the result obtained we construct the equations of mo-


tion by Chaplygin's approach. This is allowed since 'PI enters neither the
expression for the kinetic energy nor the equation for the non-holonomic
constraint. The kinetic energy of the system is

where the first term

of the expression for the reduced moment of inertia 8 (x) is obtained by


eliminating <PI by means of equation (12) for the non-holonomic constraint
from the term
1 .2
"2 8A 'PI.
Chaplygin's equations are as follows

(8.8.19)
8.8 Applications to non-holonomic systems 433

where the correcting terms calculated by eq. (7.15) are given by

d ( -R
· ·X-d
R 'P2 = 8- A'P1 R2 3 x..'P2, }
-) = - 8- A
X (x - c)
X - C
(8.8.20)
8 · · d ( -R
R x=--A'P1'P2-d -) =-A
8 .2
R2 3'P2·
X X - C (x - c)
Now it is not difficult to see that the systems of equations (19) and (17)
are identical.
The equations of motion under stationary rotation, which are the equa-
tions of relative equilibrium, are obtained by setting in eq. (17)

x=xo, x=O, x=O, <P2 = O.


We arrive at the relationships

oRo (8.8.21 )
m1--+ m 2=0,
Xo - c
Entering the quantities which define the deviations from the stationary
regime

x - Xo
- - - =q,
Xo
assuming them to be small and accounting for (21) we obtain the system
of linear differential equations

8 (xo) D. - xo<I> (xo) q. + ~R~ ~R (/11


2 .0 q = 2"70 - /12) , }
PO'P2 PO'P2 (8.8.22)
m(xo) q + ~mNrpf (q + 2D) + (C1 + C2) q = 0,

where

Let us consider the corresponding homogeneous system /11 = /12 = O.


Looking for the particular solution in the form

D = Doe At , q = qoe At ,

we obtain the characteristic equation for .x

8 (xo) m (xo).x 3+ [8 (xo) (C 1 + + ~mNrpg2) +


C2

mNxo<I> (xo) rpg2] .x + mN Xo mgrpg = o. (8.8.23)


Po
434 8. Other forms of differential equations of motion

Here m~ is replaced by means of the first equation of the relative equilibrium


(21). A dash-pot with a resisting force proportional to the coupling velocity
introduces the term with .x2 into eq. (23). Then the second equation in (22)
gains a term with q. Stability of the stationary rotation can be achieved
when power mg<,Dg > 0 is supplied to shaft 2 whereas shaft 1 acts as its
receiver, i.e. m~<,D~ < o. The rotation is stable if the coefficients of eq. (23)
satisfy Hurwitz's criterion. In the case of polynomials of the third order the
analysis reduces to Vyshnegradsky's diagram [92].

8.9 Explicit forms of the Euler-Lagrange equations


Our attention is now restricted to the case of stationary constraints and
quasi-velocities of homogeneous linear form (1.5.1) in the generalised coor-
dinates. Using expression (4.1.6) for the kinetic energy we find
aT d aT aA*
L =L +L L
n n n n
~ = A;kWk, -d ~ A;kwk !l sk Wrwk·
uWs t uWs u7r r
k=l k=l r=lk=l

Substituting this into eq. (1.8) yields


n n n n

k=l k=lt=lr=l

(s=1, ... ,n)

Applying notation (4.10.9) we can write the double sum in the form

Then, we obtain

(s=1, ... ,n).

(8.9.1)
Obviously we would arrive at the same equations by constructing Appell's
equations (6.10) with the help of the energy of acceleration (4.10.12).
Given l equations of the non-holonomic constraints

Wm =0 (m = 1, ... , l), (8.9.2)


8.9 Explicit forms of the Euler-Lagrange equations 435

equation (1) takes the form

t
k=/+l
A;kWk + t t
k=/+l t=/+l
WkWt {t l~sA;k
r=l
+ [t, k; Sl1r} = Ps
(s=l+l, ... ,n). (8.9.3)

The relationships

t
k=l+l
A;kWk + t t
k=l+l t=/+l
WkWt {t l~mA;k
r=l
+ [t, k; ml1r}
Pm + .Am (m = 1, ... , I) , (8.9.4)

in which Wk should be replaced by their expressions from eq. (3) serve to


determine of the generalised constraint forces .Am.
As an example let us determine the generalised constraint forces in the
problem of a rolling ring. It is necessary to determine forces .A4 and .A5.
The non-zero coefficients in expression (4.13.6) for the kinetic energy with
subscripts 4 and 5 are

A34 = -Ma, Ai5 = -Masin'l9.

The three-index symbols of interest are 1~4 and 1~5 which are, due to eqs.
(2.10.28) and (2.10.29), given by

5 1 4 1
124 = - --:--::a ' 1 25 - --
- sin '19.
Slnu

Analysis of formulae (4.10.9) indicates that the only non-zero bracket is

Then we obtain

(8.9.5)

The same expressions can be obtained by applying the theorem on momen-


tum. Using the "half-fixed" axes described in Sec. 2.10, we have

Taking into account


• W2
nl = - - - n
sin'19 '
436 8. Other forms of differential equations of motion

we find

or, byeq. (2.11),

The projections of the resultant vector of the external forces V on direc-


tions nand nl are the required forces.
The derivatives W3 and WI in eq. (5) should be replaced by their expres-
sion using equations of motion (2.14). Then we have

(8.9.6)

8.10 Equations of motion of a free rigid body


When constructing the Euler-Lagrange equations of motion we begin with
the expression for the kinetic energy (4.7.8)

T = ~ {M (V5l + V52 + V63) + 2M [(V02W3 - V03W2) x~+


(V03 W l - VOlW3) Y~ + (VOlW2 - v02wd z~l + 8iwi + 8;w~ + 8;wn·
(8.10.1)

Here VOl, V02, V03 and WI, W2, W3 denote projections of the velocity of the
pole 0 and the angular velocity on principal axes of inertia Ox' y' z', re-
spectively, and x~, Y~, z~ are the coordinates of the centre of inertia re-
ferring to these axes. The necessary three-index symbols are given by eqs.
(2.10.3) and (2.10.8), the quasi-velocities VOl, V02, V03 having numbers 4,5,6
by virtue of (2.10.4). Now we have

-a
VOl
aT = M (VOl + ZCW2
, ' ), }
- YC W 3
(8.10.2)
aT M
aWl =
(YC
, V 03 -
' ) 8*
ZC V 02 + - 1 WI

and the analogous equations are obtained using a cyclic permutation of the
indices.
8.10 Equations of motion of a free rigid body 437

The non-zero three-index symbols with the right subindices 1 and 4 are
2 = -1,
1'31 3=1 ,
1'21 5_
1'61 - - 1,
Turning now to the Euler-Lagrange equations we obtain

(8.10.3)

and other analogous equations. The right hand sides are written in accor-
dance with equalities (5.2.12). After the substitutions we arrive at two sets
of equations, each containing three equations.
The equations of the first set have the form

(8.10.4)

the other two being obtained using a cyclic permutation of indices. The
first equation of the second set is

e~W1 + (e; - e;) W2W3 +


M [y~ (V03 + W1V02 - W2V03) - z~ (V02 + W3V01 - W1V03)] = mf. (8.10.5)
We arrive at the same equations with the help of expression (4.11.10) for
the energy of accelerations. We rewrite it as follows

s = '12 M .2
(val + v02 . 2) + va'
·2 + v03 * [w x M ( Va + w x rc, )] + 10 6)
(8 ..

Mw· [r~ x (; +w x va)] + ~w. eO. w+ w· (w x eO. w).


Using the differentiation rules

(8.10.7)

and taking derivatives with respect to quasi-accelerations Vas and ws , we


obtain Appell's equations in the form

M {vas + (w x vO)s + (w x r~)s + [w x (w x r~U} = Vs,

M [r~ + (~O +w xva) L+ (eo. w+ w xeO . wL = m~


(8 = 1,2,3).
438 8. Other forms of differential equations of motion

Of course, these coincide with equations (4) and (5). It is clearly seen that
the equations of motion of the free rigid body are the two vectorial equalities

M [~o +w x Vo + W x r~ + w x (w x r~)] = V, (8.10.8)

eO. w + w x eO. w + Mr~ x (~o +w x vo) = mO. (8.10.9)

The first equality expresses the theorem on momentum, whilst the second
one expresses the theorem on the moment of momentum about pole O.
Indeed, recalling expression (4.8.9) for the resultant momentum Q and the
differentiation formulae (2.7.9) and (2.13.3), we obtain

(8.10.10)

that is eq. (8) expresses the theorem on momentum.


The theorem on change in the moment of momentum has the form

KO=rno ,

where the moment of momentum about fixed point (j is given by expression


(4.8.10). Taking time-derivative we obtain

KO = ro x Q + Vo x Q + M (w x r~) x Vo +
Mr~x (~o+wxvo)+eo.w+wxeo.w=rnO, (8.10.11)

where the rule of differentiation of tensor (4.3.13) and relationship (4.3.12)


are applied, to give

(eO x w) . w = eO. (w x w) = o.
The known theorem of statics and eq. (10) yields

rno = rn o + ro x V = rn o + ro x Q. (8.10.12)

On the other hand

Vo x Q + M (w x r~) x Vo
Vo x M (~·o +w x r~) + M (w x r~) x Vo =0
and eq. (11), after simplification by means of these formulae, coincides with
eq. (9) what was required to prove.
It is worthwhile mentioning that, while deriving the equations of motion
of the free rigid body from the general theorems of dynamics on momen-
tum and angular momentum, we have to overcome a number of difficulties.
8.10 Equations of motion of a free rigid body 439

These are differentiation of vectors and tensor eO referring to the moving


coordinate basis, calculation of the moment of momentum about the fixed
point (as the origin of axes fixed in the body does not coincide with the
centre of inertia) and application of the static theorem on a new centre
of moment. In contrast to this, the derivation based upon Appell's equa-
tions is formal which is typical for the methods of analytical mechanics.
The same observation is valid for the Euler-Lagrange equations, too. The
advantages of the general methods of analytical mechanics are clearer in
more challenging problems, such as multi-body systems.
Equations (8) and (9) simplify considerably if pole 0 is made coincident
with the centre of inertia C. Then rc = 0 and the equations of motion of
the free rigid body takes the form

M (;C +w x VC) = V, e C . w+ w x e C . w = mC. (8.10.13)

Let us mentally separate a part 8 1 of the body 8 using surface a. Let


Ml denote the mass of this part, rCl = CC~ and rk- = tJK the position
vectors of its centre of inertia and the point on surface a, respectively, and
efl the inertia tensor of 8 1 at its centre of inertia.
We also denote the resultant vector and the resultant moment about
point K of the active forces applied to body 8 at points on the body 8 f by
V(I) and mff), respectively, whereas those of the reaction forces acting on
surface a by Rand M K.
Taking the pole at point C we can write the equation of motion of the
body 8 1 due to eq. (8) in the form

Ml [;C +w x Vc + w x rC + w x (w x rcJ]
l = V(I) + R. (8.10.14)

Taking into account eq. (13) we obtain

R= -V(I) + ~V +Ml [w x rC +wx (w x rcJ].


l
(8.10.15)

By means of theorem (12) from statics

mg) + MCl = mff) + MK + (rk- - rcJ x (V(I) + R).


Thus, taking the centre of inertia of body 8 1 as the pole we find, by means
of the second equation in (13) and eq. (15), that

M K K
-m(l) + el
Cl ' W• + W e
X-ICl . W (8 •10. 16)

+ (rc l - rk-) x {~V + Ml [w x rC +wx (w x rcJ]}.


l

The angular acceleration vector w should be removed from eqs. (15) and
(16) with the help of eq. (13). The general expressions can be written
440 8. Other forms of differential equations of motion

by using matrix notation and introducing the inverse matrix (e C ) -1. We


restrict our consideration to the case of a body of revolution.
The vectors appearing in the equation of motion are decomposed into
components along the symmetry axis i~ and in the plane of the unit vectors
ii, i~ perpendicular to i~

Considering plane ii, i~ as the complex plane we introduce the complex


number

corresponding to vector a*. Then the complex number ia* corresponds to


the vector

lying in the same plane. Now, returning to the equations of motion (13) we
rewrite them in the form

M (~C* +i~VC3 + w* X VC* + i~W3 X VC* - vC3i~ x w*) }


= i~V3 + V*'
Aw* + Ci~W3 + (w* + i~W3) x (Aw* + Ci~W3) = i~mf + m~.
(8.10.17)

Here A and C denote the equatorial and meridional moments of inertia of


the body, so that

Separating in eq. (17) the axial and transverse components of the vectors
and using the introduced complex values we obtain two sets of equations
of motion, namely the equation for the axial motion

(8.10.18)

and equations for the transverse motion

M [vc* + i (W3VC* - VC3W*)] = V*, }


(8.10.19)
Aw* - iCC - A)W3W* = m~.

These are the equations of motion of the free body of revolution.


Proceeding to calculating the reaction force R and the reaction moment
MK we adopt that the surface a separating body Sl is the plane perpen-
dicular to the body axis at a distance Ihl from the centre of inertia S, see
8.10 Equations of motion of a free rigid body 441

I
I

FIGURE 8.3.

Fig. 8.3. The body SI is also a body of revolution about symmetry axis i~
with the centre of inertia on this axis at a distance lsi from C so that

efl = Al (i~i~ + i;i;) + CIi~i~


and

It follows from eqs. (18) and (19) that

(8.10.20)

Substituting this into eq. (15) we arrive at the expression for the axial force
in cross-section (J

R3 = -V(I)3 + MI
M V 3- MIS (2
WI +w22) (8.10.21)

and the complex-valued transverse force in this cross-section

MI MIs ( . C
R* = -V(I)* + M V* +A -2m* + CW3W* ) . (8.10.22)

Before we proceed to eq. (16) we find

AIw* + CIi~W3
Al [ A)
+ (C - C 13mC
I
A C
m* W3 13 X w*
of ]
+C *
of
,
w x efl.w (AI - CI ) w3i~ x w* '

(r~, - r~) x MI [~ + w x r~, + w x (w x r~,) ]


M 1 (s - h) (~ i~ x V + sw * + si~ x w *W3) .
442 8. Other forms of differential equations of motion

By virtue of eq. (16) we obtain the torque in cross-section a

(8.10.23)

and the complex-valued bending moment

MK
*
= K
-m(I)* +
Al + MlA(s - h) s m*G + Ml (
M s-
h) 'v,
~ *+
1
A {G [AI + Mt(s - h) s]- G1A}W3W*. (8.10.24)

The simplest example is the motion of a heavy homogeneous rigid body


with initial rotation, the gravity force being the only external force. As
m G = 0 (Euler's case), the motion of the centre of inertia of the body
is a parabolic motion of a particle in a vacuum accompanied with a free
rotation about the centre of inertia.
In this case in eq. (15)

Ml
-V(I) + M V=O,

since the weight V(I) of the separated part 8 1 is proportional to its mass.
Then

since force V(I) is applied at the centre of inertia G1 of the body 8 1 .


Equations (15) and (16) take the form

R=Ml[wxrCi+wx(wxrcJ), }
(8.10.25)
MK = efi . w+ w x efi . w+ (rci - r~) x R.

Under a pure translation R = 0 and MK = 0, as one might expect. In


this. case the mentally separated parts do not interact.
In the case of the body of revolution eqs. (21)-(24) have the form

R3 = -MIS (W~ +W~), Mf = 0, }


MIG
R* = ---::;r-SW3W*, (8.10.26)

Mf = ~ {G [AI + Ml (s - h) s]- G1A}W3W*.

The transverse force and the transverse moment become zero provided that
the body does not rotate about its longitudinal axis (W3 = 0).
8.11 Equations of motion of a spinning shell 443

FIGURE 8.4.

8.11 Equations of motion of a spinning shell


The derivation of the equations of motion assumes no wind, no rotation of
the earth and no change of the gravity force both in value and direction.
The most general expressions for the aerodynamic force and moment, eqs.
(5.13.2) and (5.13.4) , are adopted. Equations of motion of the centre of
inertia of the shell will be constructed with respect to axes of the natural
trihedron of the trajectory described in Sec. 2.18, whilst the equations of
rotation with respect to axes i~, i~, i; are bound to the shell. The unit
vector of the symmetry axis is denoted by i; (in Sec. 5.13 it was denoted
as k). Figure 8.4a shows these trihedrons, angles a and (3 are called the
angles of attack and slide, X denotes the angle of rotation of the trihedron
bound to the body about axis i; with respect to the natural trihedron.
The construction coincides with that in Sec. 2.4. Figure 8.4b shows angles
A, /-L, l/ describing the position of the natural trihedron T, n, b referring to
the fixed axes O~T]( , OT] being the upward vertical.
The relationships between the trihedrons T , n , b and i~, i~ , i3 are given
in the table below obtained under the assumption of small a and (3. This
assumption is also used in Sec. 5.13 for derivation of the aerodynamic forces.

II II T n b II
./
11 a cos X + (3 sin X cos X sinx
./
12 a sin X + (3 cos X -sinx cos X
./
13 1 a -(3

Table of direction cosines


444 8. Other forms of differential equations of motion

The angular velocity of the trihedron of axes with respect to the natural
trihedron is given by the vector

(8.11.1)

the values of the second order of smallness being neglected. As follows from
Fig. 8.4b the vector of the angular velocity of the natural trihedron is

n = T (Ii + ~sinll) + n (~COSIlCOSJ/ + IlSinJ/) +


b (-~cosllsinJ/ + j1,cosv). (8.11.2)

A comparison with formula (2.18.10) yields the relationships

R
Vc = ~n = - \. cos 11 sm
Lb A . v + 11. cos v, }

o = nn = ).. cos 11 cos J/ + 11 sin J/ , (8.11.3)


Vc n . \..
T = ~Lr = V + A sIn 11,
where R denotes the radius of curvature of the trajectory of the centre of
inertia of the shell.
The angular velocity vector w of the shell equals the geometric sum of
vectors w' and n, that is

w = T (X + i; + ).. sin 11) + n (~ + xo:) + b (a - ~ cos 11 sin v + j1, cos J/ ) •

(8.11.4)
Using the tables of the direction cosines we obtain the expression for the
axial component

W3 = X+ v + ).. sin 11 - (3 (it cos v- ).. cos 11 sin v) (8.11.5)

and the complex-valued quantity w* corresponding to the transverse com-


ponent w* of vector w

e-ixA = [i (it cos v - ).. cos 11 sin v ) +

~+ia+i((3+io:) (~sinll+i;)] e- ix . (8.11.6)

As follows from the assumptions made for deriving the aerodynamic force
and moment the value of A is small. The expression for A has the component
. (.11 cos J/
t -
\. . ) = ~R.
Acosllsmv .Vc

If the trajectory is not very steep this component has the same order of
smallness as the others and this allows the products of this component and
0: and (3 to be neglected.
8.11 Equations of motion of a spinning shell 445

We proceed to the vector of velocity of the centre of inertia. By using


the first column of the table of cosines, we have

Vc = VCT =Vc [i~ (-acosx + ;3 sin X) + i; (asinx + ;3 cos X) + i~l


and, within the adopted accuracy, we obtain

V3C = Vc (8.11. 7)

and the complex-valued quantity Vc* determining the transverse compo-


nent Vc* of the velocity vector

Vc* = VlC + iV2C = Vc (-a + i;3) e- ix . (8.11.8)

It is easy to prove that projections of vector vc* (but not of vo) on di-
rections nand b are equal to the real and imaginary parts of the complex
number vc*e ix so that

v c* = Vc (-an + ;3b), vc* = Vc (-a + i;3) . (8.11.9)


The above is valid for any vector of the first order of smallness which lies
in the transverse plane of the shell except for the vector of finite rotation
as the planes n, band ii, i~ do not coincide. Also the transverse force F *
and moment m;; are assumed to be values of the first order of smallness.
Using expressions (5.13.4) we obtain their complex-valued counterparts

F* F*e ix = pa 2 [-it vb (-a + i;3) + haw3vci (-0: + i;3) +


gla2w3A + g2iavcAJ ' (8.11.10)

me; pa 3e- ix [- f~ aW3vC (-a + i;3) - f~ivb (-a + i;3) -


g~avcA + g~iaw3Al. (8.11.11)

The vector of the gravity force is given by the expression

- M gh = - M 9 (T sin jl + n cos jl cos v - b cos jl sin v) . (8.11.12)

In order to construct the equations of motion of the centre of gravity we


should separate the coefficients of vectors T, n, b in the vectorial equation

M (TVc +n 1) = i~F3 + F* - Mgi2 = T (F3 - Mg sin jl) +


(na - b;3) F3 - Mg cos jl (ncosv - bsinv) + F*, (8.11.13)

where the real and imaginary parts of F* represent the projections of F *


on directions nand b. Recalling eq. (3), we arrive at three equations

. pa 2 h 2 .
Vc = -~vc - gsmjl, (8.11.14)
446 8. Other forms of differential equations of motion

p~
iL cos v - .x. cos f.L sin v = -Vc
9
- cos f.L cos v + (h - h) -vco: -
M

h ~ wd3 + gl ~:: [iJ - 0: ( ,\ sin f.L + v)] +


3
92~ ['\cOSf.Lsinv-iLcosv-a-;3('\Sinf.L+V)] , (8.11.15)

9 pa 2 pa 3
o - COSf.L sinv - (h -
Vc
h) -;3vc - h - W30: +
M M

91 ~~3 [a + iL cos v - ,\ cos f.L sin v + ;3 ( ,\ sin f.L + v)] +


3
92~ [iJ-o:('\sinf.L+v)]. (8.11.16)

In addition to these, we need the equation of rotation about the longitudinal


ruClS

. pa 4
W3 = -C93V3W3 (8.11.17)

and the complex-valued equation (10.19) for the transverse rotation


A (Ae- iX )· - i (C - A) W3Ae-ix = m~.

Taking into account (11) we can reduce the latter equality to the form

A - i ~ W3A + i (W3 - x) A =
3
- p~ [ff aW3vC (-0: + i;3) + f~ivb (-0: + i;3) + 9~ avc A - 9~iaw3A1.
(8.11.18)
The values of A and X are determined from relationships (6) and (5). We
have six equations of motion (14)-(18) (eq. (18) contains two equations)
and the kinematic relationship (the second equation in (3)) in the same
number of unknown variables
0:, ;3, f.L, .x, v, W3, Vc· (8.11.19)
Under such general assumptions on the aerodynamic forces, the trajec-
tory of the centre of inertia is not a plain curve since the requirement of
vanishing curvature liT given by the third equation in (3) would lead to
an increase in the number of equations with the same number of unknown
variables.
Using the less general assumption (5.13.17) on the aerodynamic forces
instead of eqs. (10) and (11) we would have

F* = F*e ix= pa 2 (fR + h) (0: - i;3) vb, }


c
m* = pa e- tX [ 3'
2
fMVC (;3 + io:) - fHavCA ,
1 (8.11.20)
8.11 Equations of motion of a spinning shell 447

where
(8.11.21)
It follows from eqs. (17) and (lO.18) that the angular velocity of rotation of
the shell about its symmetry axis retains its constant value w~. Repeating
the above calculation we arrive at the system of equations of motion of the
centre of inertia
. pa 2 1R 2 .
Vc = -~Vc - gSlllJ-l,
. g p~
it cos v - AsinvcosJ-l = - - cos J-l cos v + IL-VCO', (8.11.22)
Vc AI
g pa 2
0= -cosJ-lsinv-h-vcf3
Vc AI
as well as to the equation of rotation
. C . . pa 3 2 . pa 4 I H
A - AW3A + 2 (W3 - X) A = A IMVC (f3 + 20') - -A-vcA. (8.11.23)

The trajectory of the centre of inertia is a plain curve in the vertical


plane A = Ao under the following condition

A = Ao, v= o. (8.11.24)
In this case, the above kinematic relationship is identically satisfied and
formulae (5) and (6) simplify under the assumption that it = Vc / R, yielding

wg = X, A i
= (it + a - i~) . (8.11.25)

If the coefficient of the aerodynamic lift h is not zero then, from the third
equation (23), we have f3 = O. This is not compatible with the equation of
rotation (23) since this equation would take the form

.
J-l + 0'
. -
.cA W3 (.J-l + 0'. ) = -A--vcO'
2
pa 3 1M 2
-
pa4 1H (. .) ,
-----:;:r-vc J-l + 0'

where J-l is determined by the previous equalities and 0' is complex-valued.


Neglecting the aerodynamic lift, i.e. h = 0, simplifies the problem con-
siderably. The well-known equations of ballistics
. pa 2 1R 2 . . g
Vc = - - - V c - gSlllJ-l, J-l = - - cosJ-l (8.11.26)
AI Vc
define the motion of the centre of inertia independent of the shell rotation.
Then we have the equation for the rotation

(8.11.27)
448 8. Other forms of differential equations of motion

whose right hand side is obtained with the help of (26). Equation (27)
differs from the conventional equation of the rotational motion of the shell
only in the component proportional to the damping moment (the term with
!H), see e.g. [29].
9
Dynamics of relative motion

9.1 Differential equations of motion of a carrying


body
Let us consider a material system consisting of "carrying" rigid body and
N "carried bodies" which are particles whose position with respect to axes
Oxyz, fixed in the "carrying body" can be described by a finite or even
a countable set (the case of a solid) of the generalised coordinates. While
investigating the motion of such a system we can state two problems. The
first problem is as follows. The motion of the" carrying body" is prescribed
and the motion of the" carried bodies" is required under the assumption
that the motion of the" carried bodies" does not affect the prescribed mo-
tion of the carrying body. The position of the system can be defined by n
independent generalised coordinates qs, the case of a solid included. Par-
ticularly, such a motion can occur when the mass of the carrying body
is much greater than the masses of the carried bodies and their influence
on the motion of the carrying body can be neglected. For example, the
influence of a gyroscope on the motion of the earth can be uncondition-
ally neglected. However the motion of the earth influences the motion of
the gyroscope considerably. It is clear that this case can occur when the
prescribed motion of the carrying body is caused by external forces. These
forces can be determined from the equations of motion.
The second problem is concerned with the more general case when the
law of motion of the carrying body is unknown and needs be determined,
450 9. Dynamics of relative motion

with the influence of the carried bodies being taken into account. Generally
speaking, the system configuration is described by n + 6 parameters.
The study of motion of the system can be performed in such a way that
both cases are covered by the same approach, the first case being obtained
from the second only by omitting some equations which are conditionally
termed the equations of motion of the carrying body. This section concerns
the derivation of these equations.
Let us describe the motion of the carrying body by the velocity vector
Vo of the pole 0 and the angular velocity vector w. The position of the
"carried" particle Mi with respect to the inertial axes Oxyz and axes Oxyz
fixed in the body is described by the position vectors ri and r~, respectively.
Time is assumed not to appear explicitly in the expression for r~, i.e.

(9.1.1)

Denoting the position vector of the pole of the carrying body by ro, we
have

(9.1.2)

If the motion of the carrying body is given, then ro is a prescribed func-


tion of time and the system constraints are non-stationary since time t
appears in the expression for rio If the motion of the carrying body is to
be determined, too, then ro is determined in terms of the generalised co-
ordinates describing the position of the pole. In this case, the constraints
are stationary.
In what follows we repeatedly use the lemma of Sec. 2.13 for differen-
tiating vector a given by its projections on the axes fixed in the carrying
body

a• =a* +w x a, (9.1.3)

where i denotes the vector whose projections on these axes are equal to
the derivatives of projections a on these axes. In particular,

(9.1.4)

The kinetic energy of the system is given in Sec. 4.9 by the sum of three
components

(9.1.5)

Here the kinetic energy of the translation motion Te is as follows

Te = ~ [Mv5 +2mvo' (w x r~) +w· eO .w], (9.1.6)


9.1 Differential equations of motion of a carrying body 451

where M and eO denote the mass of the system and the tensor of inertia
at point 0, respectively.
This expression does not differ from the kinetic energy of the "frozen"
system, i.e. the system with all its bodies forming a single rigid body. Then
eO and rc would be constant. However, in our case, they are functions of
the generalised coordinates ql, ... ,qn. The value of Te does not depend on
the generalised velocities (il, ... , qn.
The next component is

(9.1.7)

where Qr and K? denote the principal vectors of the relative momenta


and the moment of the relative momenta, respectively. K? depends on the
values of va and w describing motion of the carrying body as well as on the
relative motion of the carried bodies. Tm is a linear form in the generalised
velocities with the coefficients given by equalities (4.9.8). Finally, Tr is the
kinetic energy of the relative motion and is expressed by a quadratic form
of the generalised velocities qs with the coefficients (4.9.9) determined by
the generalised coordinates.
Let us denote the projections of vectors va and w on axes Oxyz fixed
in the carrying body as VOi and Wi (i = 1,2,3, ... ). It is important to no-
tice that, due to the choice in the parameters defining the motion of the
carrying body, the kinetic energy T does not depend upon the generalised
coordinates. For this reason, we have the case (mentioned in Sec. 8.1) in
which the differential equations are split into two sets, namely a set of the
Euler-Lagrange equations for quasi-velocities VOi,Wi and a set of Lagrange's
equations for quasi-coordinates qs.
The virtual displacement of a point M of the system is, due to eq. (2),
given by

(9.1.8)

Here () stands for the vector of the infinitesimal small rotation of the car-
rying body and the equality for or~ is analogous to eq. (4). The elementary
work of all active forces applied both to the carrying and the carried bodies
due to the virtual displacement of the system is, due to eqs. (5.2.5) and
(5.1.4), given by the equality

n
O'W = V . Oro + () . rna + L Qsoqs, (9.1.9)
s=1

where V and rna denote the resultant force and the resultant moment
about pole O.
452 9. Dynamics of relative motion

Let us proceed to construct the Euler-Lagrange equations for the quasi-


velocities VOi,Wi. We have

since Tr is independent of the quasi-velocities. By virtue of eqs. (6) and


(4.7.8) we have

(9.1.10)

and accounting for eq. (4.7.4)

af)Te
Wl
= M (YCV03 - ZCV02) + egwl + egw2 + e~W3,
af)Te
W2
= M (ZCVOl - YCV03) + egwl + e~2W2 + e~3W3, (9.1.11)

= M XC V02 - YCVOl + e 31 Wl
( )
aTe
-a
W3
0
+ e 032W2 + e 033 w3.
Here Xc, Yc, Zc denote the coordinates of the centre of inertia of the whole
system referring to axes Oxyz, whilst e~ designates the components of
the inertia tensor at point O. As pointed out above, all these values should
be considered as prescribed functions of the generalised coordinates.
Using eq. (7) we obtain

aTm Q M' aTm Q M' aTm Q M' (9112)


avOl = rl = Xc, aV02 = r2 = Yc, aV03 = r3 = Zc ..

and furthermore

(9.1.13)

Expressions (10)-(13) for the derivatives should be substituted into the


Euler-Lagrange equations. The latter take the form (8.10.3) of the equations
for a free rigid body
d aT aT aT
- - - +W2-- -W3-- = Vb (9.1.14)
dt avOl aV03 aV02

d aT aT aT aT aT 0
- - - +W2-- -W3-- +V02-- -V03-- =ml, (9.1.15)
dt aWl aW2 aW3 aV03 aV02
9.1 Differential equations of motion of a carrying body 453

since only unchanged three-index symbols for quasi-velocities VOi,Wi are


used for the above construction. Additionally, the expression for T does
not contain quantities determining the position of the carrying body, thus
the derivatives of T with respect to the quasi-coordinates vanish as in the
case when constructing the equations of motion for a free rigid body.
Inserting values (10) and (12) into eq. (14) we obtain the first of the set
of three equations for the quasi-velocities

M {VOl + (W2ZC - W3YC) + (W2ZC - W3YC) + (W2 V 03 - W3V02) +


[W2 (WIYC - W2Xc) - W3 (W3XC - WIZC)] + (xc + W2ZC - W3YC)} = VI.
(9.1.16)
The other two equations are obtained by a cyclic permutation of the indices.
The vectorial form of these three equations

* +w x Vo
M [ Vo +wx r~ +w x (w x r~) + 2wx *' + **']
rc rC = V
(9.1.17)
expresses the theorem on motion of the centre of inertia or, what is equiv-
alent, the theorem on change in momentum. In accordance with Sec. 2.11,
the value in brackets presents the absolute acceleration We of the centre
of inertia of the system, with
*
We =Vo +w x Vo + w.x,
rC +w x (
w x r ,C ) (9.1.18)
being the translational acceleration composed of the acceleration of the
pole
* +w x Vo,
Wo =Vo (9.1.19)
and centripetal and rotational accelerations

(9.1.20)
respectively.
Since the axes are rigidly bound to the carrying body their angular ve-
locity coincides with that of the carrying body and thus
*
w. =W* +w x w =w. (9.1.21)
The term in eq. (17)

(9.1.22)

is the Coriolis acceleration, r*'c denoting the relative velocity of the centre
of inertia of the system. Finally

Wr
**'c
=r (9.1.23)
454 9. Dynamics of relative motion

is the relative acceleration. Calculation of the relative velocity and relative


acceleration is performed by means of formulae (2.14.5)

(9.1.24)

We proceed now to the second set of equations of motion of the rigid


body. Inserting eqs. (10) and (12) into eq. (15) and simplifying the result
we obtain

M + eg WI + egW2 + egW3 + WI eg + W2eg +


(YCV03 - ZCV02)
.° + Krl
W3e13 .° + M [W2 (V02XC - VOlYC) - W3 (ZCVOI - XCV03)] +
W2 (e~1 WI + e~2W2 + e~3W3) - W3 (eg WI + egW2 + e~3W3) +

w2K~ - w3K~ + M [V02 (WIYC - W2Xc) - V03 (W3XC - WIZC)] = m?


(9.1.25)

The corresponding vectorial form is as follows

* ° ·W + w x eO . w+ K* ° +
eO . w +e r

W x K? + Mr~ x (~O +w x Vo) = mO. (9.1.26)

In order to transform this equation we use definition (4.3.20) of the tensor


of inertia eO to obtain

(9.1.27)

Accounting for eq. (4.9.6) we can write

*0
e ·w+w x K?

or, after simplification,

*0
e ·w+w xK?

(9.1.28)
9.1 Differential equations of motion of a carrying body 455

Placing these terms in the right hand side of eq. (26) we can treat them as
the moment of the Coriolis forces of inertia

°
mCor = - e*0 ·W - w x Kr . ° (9.1.29)
This moment is applied to the carrying body and appears as a result of the
relative motion of the carried bodies.
By using eq. (17) we can easily remove the acceleration of the pole from
eq. (26). The result is
*C
eC . w + w x eC . w = mC + mccor - Kr (9.1.30)
Here e C denotes the moment of inertia of the system at its centre of inertia
(9.1.31)
The resultant moment of relative momenta about the centre of inertia is
(9.1.32)
and the moment of the Coriolis forces of inertia about the centre of inertia
is

(9.1.33)
Finally, m C denotes the resultant moment of the active forces about the
centre of inertia
(9.1.34)
Of course, eq. (30) can be obtained directly from eq. (26) if pole 0 is
coincident with the centre of inertia C of the system.
We considered the first part of the problem, that is, we derived the system
of equations of motion (17) and (26) or (17) and (30) for the carrying body.
This system is not complete as it contains the generalised coordinates
qs as parameters, thus the equations for relative motion of the carried
bodies should be added. Obviously, when relative motion is absent, eqs. (17)
and (26) coincide with the equations of motion of the rigid body (8.10.8),
(8.10.9).
We note in passing that if the carrying body has a point moving in a
straight line with constant velocity with respect to an inertial coordinate
system O~'T}( then the acceleration of the pole 0 (19) vanishes and the
carrying body can be viewed as rotating about the motionless point o. Its
equations of motion are
*0
eO. w+ w x eO . w+ Kr = m O + m8or, (9.1.35)
and eq. (7) serves to determine the constraint force at point 0, i.e. the
force to be applied in order to ensure the presence of a point with zero
acceleration.
456 9. Dynamics of relative motion

9.2 Differential equations of the relative motion of


carried bodies
Required are Lagrange's equations for coordinates q1, ... , qn determining
the position of the carried bodies mounted on the carrying body. Due to
eqs. (1.5) and (1.9), these equations are
(9.2.1)

We begin with calculation of the first component. The kinetic energy of the
translational motion Te does not depend on the generalised velocities qs,
thus
aTe arc 108°
Es(Te) = -- = -M(vo x w)· - - - - - ·w. (9.2.2)
aqs aqs 2 aqs
We proceed now to the second component. Using eqs. (1.7) and (1.8) and
taking into account that vector va does not depend on qs and qs we obtain

By virtue of eqs. (1.3.5), (1.3.11) and (1.3) we have

(9.2.3)

and thus

Hence,

* ) arc arc * arc


Es (va· Qr) = M ( Va +w x Va . ~
uqs
+M (va x w)· ~
uqs
= M Va . - .
aqs
(9.2.4)
Vector w does not depend on qs and qs, too, therefore we obtain by analogy

But

w· (w x aKO)
aq~ = 0
9.2 Differential equations of the relative motion of carried bodies 457

and the obtained expression can be written in the form

(9.2.5)

where an asterisk implies that the differentiation with respect to time is


taken with respect to axes bound to the carrying body.
The differential equations (1) are now cast in the form

*
Qs - M ( Vo +w x Vo ) . ~ arc +
uqs
1 oeo
-w . - - . w -
oKo
W . __ T - W . £* (KO) (9.2.6)
2 oqs oris ST·

We turn our attention to the terms on the right hand side of this equation.
The vector

SO = -Mvo = -M (~o +w xvo) (9.2.7)

applied at the centre of inertia of the system is called the force of inertia
for pure translation. Then, due to eq. (5.1.3), the value

arc-
So . - = -M Vo (* +w x Vo ) .arc
--
oqs oqs

a
--;;-M Vo
uqs
(* +w x Vo )'
. rc (9.2.8)

is termed as the s - th generalised force of inertia for pure translation,


whereas the scalar product

IIo = M Vo (* +w x Vo )'
. rc (9.2.9)

can be viewed as the potential energy of the homogeneous field of these


forces.
The quadratic form in projections of the angular velocity vector

IIW = - ~w . eO . w (9.2.10)
2
can be referred to as the potential energy of the centrifugal forces and the
quantities

(9.2.11)

are named the generalised centrifugal forces.


458 9. Dynamics of relative motion

We proceed to the last two terms on the right hand side of eq. (6) which
do not have the character of potential forces. Recalling the following ex-
pression

(9.2.12)

we obtain

• U
~KO
w·--=w·
~.
r • L m ·'r · x - = L m· (wxr·
N
t t
~ ,N
uri
~
. ') . - t "
~ ,
uri
~ .
uqs i=l uqs i=l uqs

The vector

(9.2.13)

represents the rotational force of the inertia of the particle. Thus, the quan-
tity

(9.2.14)

is the s - th generalised rotational force of inertia.


Returning to expression (12) we have

so that

n N "'" "'"
-w·vc*s (KO)
r
" , . ' " m iuri
=-2w·LqkL uri
- x -. (9.2.15)
s=l i=l aqk aqs

This expression can be represented in either of two forms: the first form

-w· E; (K?)

(9.2.16)
9.3 Relative equilibrium 459

displaying that the considered component on the right hand side of the
differential equation (6) is the generalised Coriolis force of inertia. Second,
recalling Sec. 7.3 we can treat expression (15) as the generalised gyroscopic
force
n
-w· £; (K~) = fs = L 'Yskqk (9.2.17)
k=l

with the gyroscopic coefficients (7.3.6)

(9.2.18)

Using the introduced notation, the system of differential equations for


the carried bodies is written in the form

(9.2.19)

The left hand sides of these equations depend only on the quantities de-
scribing the configuration and motion of the carried bodies with respect to
the carrying body. The equations of the relative motion gain the form of
equations of the absolute motion by placing the terms caused by motion of
the carrying body and the above forces of inertia on the right hand side.
If the motion of the carrying body is not prescribed, the equations in
(19) should be considered together with the equations of motion (1.17) and
(1.26). We obtain the system of n + 6 differential equations of the second
order for the generalised coordinates ql, ... , qn and six first order differential
equations for the quasi-velocities VOk and Wk. The six equations (1.5.8)
which relate the generalised coordinates corresponding to the generalised
coordinates of the carrying body to the quasi-velocities should be added to
this system of equations.
When the motion of the carrying body is given, then only equations
(19) should be considered. In this case there is no need to make use only
of this form of equations. It is necessary to have the generalised forces
of inertia on the right hand sides of the above equations. Accounting for
these forces properly we can write the equations of motion in terms of the
quasi-velocities, to use Appell's equations etc.

9.3 Relative equilibrium


When considering the relative equilibrium of carried bodies determined by
the equalities

qs = q~ = const (8 = 1, ... , n), (9.3.1)


460 9. Dynamics of relative motion

we come to the following natural questions: i) under what motion of the


carrying body is the relative equilibrium feasible, ii) how to determine
the possible relative equilibria and iii) are the positions of the relative
equilibria stable or unstable. The latter question in its general form is
beyond the scope of the present book. Answering the first question we
restrict our consideration to such motions of the carrying body under which
the acceleration of any point remains constant in its value and direction
with respect to axes fixed in the body. The formula for accelerations in a
rigid body (2.11.1) says that it occurs under the constant vectors of the
pole acceleration Wo and the angular velocity w. The generalised forces Qs
in eq. (2.19) are assumed not to depend explicitly on time.
Under the relative equilibrium of the carried bodies, the generalised rel-
ative velocities qs and accelerations qs equal zero. Thus both the kinetic
energy of the relative motion Tr and the gyroscopic forces r s become zero.
In addition to this, according to the adopted assumptions on the motion of
the carrying body, the generalised rotational forces of inertia Q~ are equal
to zero also.
As follows from eq. (2.19), the generalised coordinates q~ satisfy the
equations

(9.3.2)

where, if Qs depends on the generalised velocities qs, the latter should be


replaced by zeroes. In eq. (2)

(9.3.3)

Vector rc and the tensor of inertia eO depend on the required coordinates


q~.
If the system of n equations (2) has a solution (1) (one or several), it
determines the position of relative equilibrium. The stability of the obtained
positions of relative equilibrium is studied by the general methods of the
theory of the stability of motion for each of the possible positions of relative
equilibrium separately.
The above can be generalised. Let the coordinates describing the position
of the carried bodies relative to the carrying body include the positional
coordinates ql, ... , qm and the cyclic coordinates qm+l, ... , qn. Then equation
(2.19) takes the form

Em+I(Tr ) =r m+l +Q*'+l (l=l, ... ,n-m), (9.3.5)


9.3 Relative equilibrium 461

where Qs, II, IIo, IIw, as well as the gyroscopic coefficients and coefficients
ask of Tr depend only on the positional coordinates. This follows from
the definition of the cyclic coordinate and from the fact that eq. (2.19) is
nothing else than another form of the differential equations of motion

Provided that Wo and ware constant, the motion under which both po-
sitional coordinates

qs=q~ (s=l, ... ,m) (9.3.6)


and the cyclic genemlised velocities

qm+l = q!+l (I = 1, ... , n - m) . (9.3.7)

remain constant is referred to as the relative equilibrium. It follows from


expression (7.4.6) for the Eulerian operator
n n n
cs(Tr) = I>Skiik+ LL[k,j;slqkqj
k=1 k=1j=1
and the definition of Christoffel's symbols, that for the positional coordi-
nates
n-mn-m
L L [m + l, m + r; sl qm+lqm+r (9.3.8)
1=1 r=1

(s = 1, ... ,m),

whereas for the cyclic ones


n-mn-m
LL [m+l,m+r;m+tlqm+lqm+r =0
1=1 r=1
(t = 1, ... , n - m) (9.3.9)
Recalling now expression (2.17) for the gyroscopic forces we can recast
equations for the relative equilibrium (4) and (5) in the form

n-m
L 'Ys,m+lqm+l = 0 (s = 1, ... , m), (9.3.10)
1=1
462 9. Dynamics of relative motion

FIGURE 9.1.
n-m
2:: 'Ym+l ,mHQmH = 0 (l = 1, ... , n - m). (9.3.11)
k=l

The skew-symmetric determinant of the system of equations (11)

0 'Ym+l,m+2 'Ym+l ,n
'Ym +2,m+l 0 'Ym +2 ,n
Dn - m = (9.3.12)

'Yn,m+l 'Yn ,m+2 'Ynn

is zero when n - m is odd. Then the system of homogeneous equations


(11) has non-zero solutions for cyclic generalised coordinates and thus the
adopted definition of the relative equilibrium makes sense. In general, de-
terminant (12) is not zero when the number of cyclic coordinates is even,
so that the introduced concept is meaningful only for such values of the
gyroscopic coefficients which ensure that D n - m = O.

9.4 Equilibrium of rotating flexible shaft


We consider an elastic rod with round cross-section whose end 0 rotates
with constant angular velocity w about axis Oz. The rod is clamped into a
rigid body S at end M, see Fig. 9.1. When the mass of the rod is neglected,
the system configuration is described by the coordinates and the angles of
rotation of the cross-section M with respect to the axes fixed in the carrying
body O. It is assumed that this configuration remains under the relative
equilibrium.
9.4 Equilibrium of rotating flexible shaft 463

The coordinate axes M x' y' Zl are fixed in the body, axis M Zl being di-
rected along the tangent to the elastic axis of the rod at point M. The
coordinates x, y of point M referring to the system of rotating axes Oxyz
are denoted as u, v . As the rod is assumed to be inextensible, the vertical
coordinate of point Mis z = t. When the rod does not rotate, the rod axis
lies along axis Oz whilst axes MXly'ZI occupy the position MoxoYozo being
parallel to axes Oxyz. Let 8 denote the vector of rotation under which axes
MoxoYozo become parallel to axes M x' y' Z'. We also adopt the assumption
and notation of Sec. 5.9, that u, v and the projections Q, (3, "1 of the rotation
vector are small, and is and i~ denote the unit vectors of axes Oxyz and
M x' y' Zl, respectively.
The potential energy of elastic forces II and centrifugal forces IIw should
be expressed in terms of the five introduced generalised coordinates. We
start with the latter one. In expression (2.10) for IIW we replace the tensor
of inertia at point 0 in terms of the tensor of inertia at point M by means
ofeq. (4.4.1)

eM +M(ErM.rM-rMrM)+

2M [ErM . r~ - ~ (rMr~ + r~rM)] (9.4.1)

-=---+
Here r M = OM denotes the position vector of point M with the origin at
point 0, rc = Me
the position vector of the centre of inertia with the
origin at point M, then

(9.4.2)

For further analysis we need the table of direction cosines of the angles
between the unit vectors of trihedrons is and i~. The value of IIW should be
calculated with accuracy up to the squares of the generalised coordinates.
This forces us to retain the second powers of small quantities Q, (3, "1 in the
expressions for cosines. Using Rodrigues's formula (3.2.1) we have

·1·
Is = Is + 11 (8.x + - 88 . .
Is
1 Is -
1.
-21s
(}2) ,
1 + _(P 2
4
and the table of cosines within the mentioned accuracy becomes

il h i3
.,
11 1-
1
"2 ((32 + "12)
1
"1 + "2 Q(3
1
-(3 + -Q'Y
2
.,
12 -"1
1
+ -Q(3
1
1 - "2 (72 + (2)
1
Q + "2(3"1
2
., 1 1
-Q + -(3"1 1 (2
Q +(32)
13 (3 + "2Q'Y 1-"2
2
464 9. Dynamics of relative motion

Replacing vector w by wi 3 , due to eq. (2.10), we obtain

II w =
1 2·13' 80
--w .
- . 13·
2
We have

. [E
13' I
rM' rc -"2 1( I
rMrc I
+ rcrM )]. I '
. 13 = rM . rc I
- 13' rCrM •
. 13

= (Ui1 + vi2) . r~ + li3 . r~ - i3 . r~l = (Ui1 + vi2) . r~.


As one might expect the terms with l canceled out. Since u and v are small,
we can keep only linear components in the table of cosines. Then we have

( Ul1
• V12 ) . r Ic = uXc
+• I + vYc
I + (xcv
' - Ycu '({3
' ) + zc u - va ) .

Let 8 ik denote the components of tensor 8 M referring to the axes


M x' y' Zl fixed in the body, then with the help of eq. (4.4.11) we obtain

8 11 ai3 + 822a~3 + +8 33 0053

281200130023 + 282300330023 + 28 130013 0033

or, by means of the table of cosines,

(8 11 - 8 33 ) {32 + (8 22 - 8 33 ) 00 2 + 8 33 -

28 12 a{3 + 28 23 (00+ ~(3/, ) + 28 13 ( -(3 + ~a/') .


Now by virtue of eq. (2.10) the potential energy of centrifugal forces is as
follows

II W _~W2 [(8 11 - 8 33 ) {32 + (8 22 - 8 33 ) 00 2 - 28 12 a{3 + (8 23 {3+

e 13 ah + M (u 2+ v 2 ) + 2Mz~ (u{3 - va) + 2M/, (x~v - y~u)]

_w 2 [8 23 00 - 8 13 {3 + M (x~u - y~v)l- ~833W2, (9.4.3)

where the latter terms can be omitted since it does not depend on the
generalised coordinates. We would loose essential terms in the expression
for the potential energy if we kept only linear components in the expressions
for cosines.
The formula for the potential energy IIe of the elastic strain of bending
and twisted rod is derived in Sec. 5.9. Let us mentally apply forces V,?, VyO
and the bending moments Lg, L~ to the end M of the rod as shown in
9.4 Equilibrium of rotating flexible shaft 465

FIGURE 9.2.

Fig. 9.2. The bending moments Ly and Lx at cross-section z of the rod are
given by

(9.4.4)

and the potential energy of bending in planes zx and yz is as follows

The coefficients of these quadratic forms form matrices of influence. The


inverse matrices, which are the matrices of stiffness, are used to construct
the expression for the potential energy of elastic forces in terms of the
generalised coordinates

IIe =
112E1
2'-[3-
[2 2
U +v -
[
22' (u,B - vo:)
[2 (2
+ '3 0: +,B
2)] + 2'[1'.
1C 2
(9.4.5)

The latter term corresponds to the potential energy of torsion. Neglecting


the potential energy of weight we arrive at the equations of equilibrium
466 9. Dynamics of relative motion

which, in the expanded form, take the form

U( M[3w
T 12EI
2 -1
)
- (3
(6EI
M[3w2 +T
Z~) y~
+ -z '"Y =
x~
-Z-,
v ( 12EI
y M[3w2-1
)
+a
(6EI
M[3w 2 + f
Z') -f'"Y=f,
x' y'

(3 ( 4EI _ 8 11 - 8 33 ) 8 12 _
M[3w2 M[2 + M[2 a
U(6EI Z~) 1 8 23 8 13
T MZ3 w2 + -Z - "2 M[2 '"Y = - M[2 , (9.4.6)

4EI 8 22 - 833) 8 12 (3
a ( M[3w2 - M[2 + M[2 +
v (6EI Z~) 1 8 13 823
Y M[3w2 + T -"2 MZ2 '"Y = MZ2'
U y~ v x~ 8 23 8 13 C
T-z -Y-z - 2MZ2(3 - 2M[2a + M[3 w2'"Y = o.
In what follows we restrict our consideration to the case when the body
mounted on the elastic shaft is a body of revolution with axis M z' being
its axis of symmetry. Then

and the latter equation says that '"Y = 0, i.e. the rod is not twisted. We
introduce the notation
3EI 2 z~ _ r 8 11 - 8 33 1 .
Ml3 w 2 = V , T - '>C, Ml2 =k, T(U+W)=P' a+i(3=c.
(9.4.7)

We notice that the case v = 1 corresponds to the angular velocity being


equal to the natural frequency of the mass M on the elastic rod. If param-
eter k is positive, the body is said to be a cylinder and if k is negative, then
the body is called a disc. For example, for the homogenous solid cylinder
k < 0 for h < rV3 and k > 0 for h > rV3 where rand h denote the radius
and the height of the cylinder, respectively.
We obtain the system of homogeneous equations

(9.4.8)

with the determinant

~ (v 2) = (4v 2 - 1) (~v2 - k) - (2v 2 + (c) 2 = ~ (v 4 - av 2 + b) ,


(9.4.9)
9.4 Equilibrium of rotating flexible shaft 467

where

(9.4.10)

If the determinant (9) is not zero, then, in the equilibrium position c: =


p = 0 the axis of the rod is a straight line. The angular velocities which are
roots of equation ~ (v 2 ) = 0 are referred to as the critical velocities. At
the critical velocities the system of equations (8) has solutions determined
up to an arbitrary factor. Then there exist a series of the equilibrium forms
close to a straight line. Sylvester's criterion ensuring positive definiteness
of the quadratic form II + IIw reduces to the single inequality ~ (v 2 ) > o.
Based on the Lagrange-Dirichlet theorem we can state that the straight
form of equilibrium is stable when this inequality holds true and it is no
longer stable when the inequality has an opposite sense.
The determinant (9) can be cast in the form

a2 - 4b = 9[(:c + ~ + k) ~ (k - (2)]
2 - (9.4.11)

3[4 ((c + ~) + 6((c + ~) (k - ~) + 3(k - ~ r].


2

It presents the quadratic form in (c + ~ and k - ~ and since 4·3 - 32 > 0

this quadratic form is positive definite becoming zero only when (c = -~


1 .
and k = "3. The roots of equatIOn ~ (v 2 ) = 0 are always real-valued. Only
one root is positive if

b< 0 or (2) k. (9.4.12)

When

b > 0, a > 0 or k > (2, (9.4.13)

both roots v~ and v~ are positive. The case of two negative roots would
take place under the condition

b> 0, a < 0, (9.4.14)

which is not feasible. Indeed, if these inequalities held, the second inequality
could be strengthened by replacing k with a smaller quantity (2.
It results
in the equality
468 9. Dynamics of relative motion

which does not hold for real values of (c.


Summarising we can state that

(9.4.15)

under condition (12), whereas

(9.4.16)

under condition (13).


The case of a single critical velocity wi occurs for any disc since b < at
k < 0. This case describes the influence of the gyroscopic effect on the
°
critical angular velocity of the disc, [6] and [31]. If we take k =
(c = 0, i.e. a particle on the shaft is considered then a = 1, b = and as°°
and

relationships (7) and (9) show, the critical angular velocity Wo is equal to
the frequency of free oscillation of this mass. For a thin disc we can take
(c = 0, then

and as k < °it is easy to prove that the inequality


1
2< lI i<1
holds (it is necessary to take into account that y'9k 2 + ... ~ -3k for k ---->
-00). Thus,

wi
2>2"=2">l.
1
(9.4.17)
Wo VI

The gyroscopic effect increases the critical angular velocity Wo of the thin
disc. Strictly speaking the word" gyroscopic" is contrary to the essence of
the phenomenon since the centrifugal force causes the instability.
Two critical angular velocities wi and w~ corresponding to the roots vi
and v~ occur under condition (13), that is for "cylinders". If v~ denote the
smaller root of (9) then, as eq. (16) shows, the straight form of equilibrium
is stable if w 2 < wi or w 2 > w~ and is unstable if w 2 < wi < w~.

9.5 A gyroscope in Cardan's suspension mounted


on a moving platform
We introduce three trihedrons of unit vectors: e s , is, i~ bound to the plat-
form, outer ring and inner gimbal, respectively. The rotor rotating with
9.5 A gyroscope in Cardan's suspension mounted on a moving platform 469

f'J
HS O.g.

FIGURE 9.3.

angular velocity rp, is mounted on the inner gimbal. The bearings of the
axis of rotation of the outer ring lie on axis e3 of the platform and those
of the gimbal on axis i2 of the outer ring, so that e3 = h, h = i~. The
axis of the rotor rotation coincides with the unit vector i~, see Fig. 9.3.
The origins of all axes coincide and are placed at point 0 where the axes
of rotation of the outer ring, the gimbal and the rotor intersect. This point
is the centre of inertia of the outer ring and the rotor. An additional point
mass is attached to the point with the radius vector
(9.5.1)
The acceleration of point 0 which is taken to be constant both in value
and direction is denoted by Wo and the constant vector of the angular
velocity of the platform by w. In order to take into account the force of
gravity, we enter the vector of the geometric sum

w* = Wo - g, (9.5.2)
with g being the acceleration due to the gravity force. Then, we set in eq.
(3.10) 11* = 119 + 110
11* = mw* . r', (9.5.3)
119 and 110 being the potential energy of the weight and the inertial forces
for pure translation, respectively.
In order to simplify and reduce the calculation of the derivatives with
respect to the generalised coordinates, it is necessary to use relationship
(1.3.5)

In our case
r=ro+r', v=vo+wxr'+r*'.
470 9. Dynamics of relative motion

As ro does not depend on qs, whereas Vo, w, r' do not depend on qs, we
have

or*' (9.5.4)
oqs .
The relative velocity r*' of the point of the rigid body rotating about
point 0 with angular velocity w' relative to the platform is w' x r', thus

or' *'
araw'
-=-=-xr' (9.5.5)
aqs oqs oqs '
as r' does not depend on qs. The vector w' depends linearly on the gener-
alised coordinates qs whilst the dependence of r' on the generalised coor-
dinates is given by more complicated transformation formulae expressing
the coordinates of the point with respect to axes bounded to the plat-
form and the body. This is the explanation for the simplifications given by
relationships (5).
Let 0: and (3 denote the angles of rotation of the outer ring relative to
the platform and the gimbal relative to the outer ring, respectively. Then
according to Fig. 9.3 we obtain the following tables of cosines

II II el I e2 I e3 II II ., II el e2 e3 II
II h II coso: I sin I 0 0: II II 11 II cos (3 cos 0: I cos (3 sin I - sin (3
0: II
II h II -sino: I coso: I 0 II II i~ II - sino: I coso: I 0 II
II i3 II 0 I 0 I 1 II II i~ II sin (3 cos 0: I sin (3 sin 0: I cos (3 II
as well as the following expressions for the vectors of the relative angular
velocities of the outer ring, the rotor and the gimbal

(9.5.6)

respectively.
We proceed now to construct equations (3.10) and begin by calculating
the derivatives of II* with respect to 0: and (3. By virtue of eqs. (39, (5)
and (1) we obtain
9.5 A gyroscope in Cardan's suspension mounted on a moving platform 471

Denoting the projections of w* on the platform axes by Ws we have

then, by means of the tables of the direction cosines, we obtain the rela-
tionships

aII*
aa = -mwI (
al( 3 sm
cos . a + a2 cos a + a3 sm. (3.sm a ) +
mW2 (al cos (3 cos a - a2 sin a + a3 sin (3 cos a) ,
(9.5.7)
aII*
a(3 = -mal ( . (3 cosa+W2 sm
Wlsm . (3 sma+W3cos
. (3) +

ma3 (WI cos (3 cos a + W2 cos (3 sin a - W3 sin (3) .

We turn our attention to construction of the potential energy of the cen-


trifugal forces. The tensors of inertia at point 0 of the outer ring, the
gimbal with the additional mass and the rotor are given respectively by

e? = GIili l + BIi2h + AIi3i3,


e?I = [BII + m (a~ + aD] i~i~ + [AIl + m (ai + a§)] i~i~+
[GIl + m (a§ + a~)] i~i~ - m [ala2 (i~i~ + i~iD + (9.5.8)
a2a3 (i~i~ + i~i~) + a3al (i~i~ + i~ i~)l ,
e?1I = AIl I (i~i~ + i~i~) + GII Ii~ i~.

Here A and B denote the moments of inertia about the central axes lying
in the mid-plate of each part and G about the axes perpendicular to this
plane. As a result, the expression for IIW takes the form

(9.5.9)

where the constants A, B, G can be easily expressed in terms of the values


entered by formulae (8). The differentiation of scalar products w· i: is
performed by means of rule (5), i.e.
472 9. Dynamics of relative motion

As follows from eq. (3.10) we are interested only in the coefficient of in tj;2
the expression for the kinetic energy Tr of the relative motion. This coeffi-
cient is constant and equal to !CIII and thus the corresponding component
in eq. (3.10) drops out. In order to calculate the gyroscopic coefficients 'Yacp
and 'Y{3cp' we notice that in this particular case eq. (3.11) is an identity, i.e.
'Ycpcp = O. According to eq. (2.18) the following expressions

are required to be found where the summation should be performed over


all particles of the system. Thrning to eq. (5) we have

ar~
- ' x-'
ar~
aw' x r' ) x (awl
( ---1!.L ---1!.L x r~ )
aa acp aCt ' atj; ,
-r~, [(awaa1II x r~)
' . aWIII] , , . (awaCt1II x awatj;III )
atj; = r~r~
and moreover

'Yacp = 2w· t mir~r~ (a~~I


i=l
. x a~~II),
cp
)
(9.5.11)
(awl aWl)
= 2w. '""' m·r/.r~· -...!,lL x ---1!.L .
N
'V
I {3cp ~, , , a(3 a.
~1 cp

Now, recalling definition (4.3.20) of the tensor of inertia we find

N N 1
Lmir~r~ = ELmir~. r~ - eO ="2E (eft + e~2 + e~3) - eO
i=l i=l
(9.5.12)

and particularly, for the rotor which is a symmetric body,

~E (2AIII + CIII) - [AIIIE+ (CIII - AIII ) i~i~l

~ECIII + (AIII - CIII) i~ i~. (9.5.13)

Thus we have

'Yacp = [CIIIW + 2 (AIII - CIII) W . i~ ii] . (e3 x iD = CIIIW . i~ cos (3, }


= [CIIIW + 2 (AIII - CIII ) W . i~ i~] . (i~ x i~) = -CIIIW . i~
'Y{3cp
(9.5.14)
9.5 A gyroscope in Cardan's suspension mounted on a moving platform 473

and furthermore

ra = GIl[cp (-WI sin a +W2 cos a) cos 13, }


(9.5.15)
r,6 = -GIl [cp (WI sin 13 cos a + W2 sin 13 sin a + W3 cos 13) .
These results can be anticipated since expressions (15) obtained are noth-
ing else than the projections of the vector of the gyroscopic moment

r=GIl[cpi~xw, r a =r·e3, r,6=r.i~ (9.5.16)

on axes of rotation of the outer and inner rings.


Now everything is prepared for constructing equations (3.10) describing
angles a and 13 under relative equilibrium. Let us consider the problem of
the influence of the ship motion on the spherical earth and the angular
velocity of the earth on the equilibrium position of the rotor in Cardan's
suspension. The unit vectors of the trihedron eI, e2, e3 are taken as being
directed along the axes of the geocentric system: el pointing north along
the tangent to the meridian, e2 pointing west along the tangent to the
parallel circle, and e3 along the upward local vertical. Then, denoting the
northern and eastern components of the vector of the platform velocity as
VN and vo, respectively, and the latitude as <P we have

VO) . VN
W = ( R cos <P + U (el cos <P + e3 sm <p) + e2 If· (9.5.17)

Our consideration is limited to taking into account terms proportional to


the first order of projections of this vector and, for this reason, the cen-
trifugal forces of inertia are excluded from the equilibrium equations.
The projections of vector w* which is the geometrical difference of the
platform acceleration and the gravity acceleration, are determined by means
of formulae (2.15.8). In these formulae the terms depending on the vertical
component of the velocity Vh and terms proportional to U 2 are omitted.
The values of velocities VN and Vo are taken to be constant. We obtain

WI = .ft tan <P + 2voU sin <P,


v2

VNVO .
W2 = ~tan<p + 2vNUsm<P, (9.5.18)
v 2 +v 2
W3=g- N R 0 -2voUcos<P.

These expressions are actually some functions of time because of presence


of the latitude. With this in view, the present consideration of the problem
on relative equilibrium is acceptable only in the case of slow change in
latitude and during short intervals of time.
Let us take
474 9. Dynamics of relative motion

in eq. (7) which corresponds to a special case of the mass in the gyrocom-
pass. We arrive at the following system of equations
1
- (WI.
sma - W2
.
cos a) sm(3 = -
(WI.
sma - -W2 cos a ) cos (3,
-
9 U U
-1 (WI cos a + W2 sma
. SIn (3
) cos (3 - -W3.
9 9
= q [(u:; a
cos + ~ sin sin (3 + a) cos (3] , 7J
(9.5.19)
where

(9.5.20)
mga
The latter two equations allow us to determine the deviation a and (3 of
the rotor axes. Since they are small, we neglect their square values, to get

(9.5.21 )

The coefficients of these equations contain components of different orders


of smallness. The values representing the ratios of the northern and eastern
component of the ship velocity to the local velocity of the rotating earth
Va
(9.5.22)
RU cos <I> , RU cos <I> ,
respectively, are small compared to unity.
The ratios of the components of the Corio lis acceleration to 9
2v N U cos <I> 2vaU cos <I>
(9.5.23)
9 9
are values of higher orders of smallness than those in (22). To prove this we
construct the ratio of terms in eq. (22) to those in eq. (23). It is proportional
to a small parameter RU 2 / 9 which is the ratio of the centrifugal force on
the equator to the gravity force.
Keeping only the principal terms in the equilibrium equations (21) we
arrive at the formulae
VN
a = (3 = -qsin <I>, (9.5.24)
RU cos <I> ,
describing the statical deviations which are the deviation of the rotor axis
to the west caused by the northern component of velocity and the deviation
over the horizon. If we account for the values of order (22) then
a = __V_N_ _ _ _ _ _1..".,...__ ~ VN (1 va) (9.5.25)
RU cos <I> 1 + Va ~ RU cos <I> - RU cos <I> '
RU cos <I>
9.6 Relative motion of rigid bodies 475

1+ Va
f3 = -q sin <I> 1 +~~~~~<I> = -q sin <I> (1 + RUV::Os <I> - q cos <I> ) •
(9.5.26)
However it would be inconsistent to look for higher order approximations,
as the terms depending on the centrifugal forces were omitted during the
derivation of eq. (21). In addition to this, the step from eq. (19) to eq. (21)
would need a further refinement.

9.6 Relative motion of rigid bodies


As a particular case of the general equations obtained in Secs. 9.1 and 9.2
we consider the case of carried rigid bodies. The analysis will be limited
to the case of a single rigid body since the generalisation for an arbitrary
number of bodies is trivial. In other words, we study the problem of the
relative motion of two bodies: one carrying body and one carried body.
Let us remember that the axes Oxyz with the pole at point 0 are bound
to the carrying body specified by index 1. The origin of the axes C 2 x' y' z',
fixed to the carried body marked by index 2, is taken at the centre of inertia
C 2 . Thus,

(9.6.1)
where r~ denotes the position vector of the centre of inertia of the system
introduced byeq. (1.17), whilst r~, and r~2 designate the position vectors
of the centre of inertia of the carrying and carried bodies, respectively.
The generalised coordinates of the carrying body are the three quantities
ql, q2, q3 describing the position of the centre of inertia with respect to axes
Oxyz and three angular coordinates q4, q5, q6 describing the orientation of
the trihedron C 2 x'y'z' referring to axes Oxyz. Hence

(9.6.2)

(k = 1,2,3), (9.6.3)
with i~ being the unit vectors of the axes of the carried body.
Since the position vector r~, is a constant vector relative to axes Oxyz,
we have the following obvious equalities

*' M2 r*'C2 ,
M rc= M **' **'
rc= M2 r C2 , (9.6.4)

(9.6.5)
476 9. Dynamics of relative motion

The equation of motion (1.17) of the centre of inertia of the system of


bodies has the form

M1 [~a +W x Va + W x r~l + WX (W x r~J] +


M2 [~a +w x Va + W x r~2 + W x (W x r~J + 2wx ;~2 + *;~2] = v.
(9.6.6)

Along with Va and w, this equation contains the generalised coordinates


Q1, Q2, Q3and the corresponding generalised velocities and generalised ac-
celerations.
We proceed to construct the second vectorial equation, which is the
"equation of rotation" (1.26). First of all it is necessary to notice the equal-
ity

(9.6.7)
where 9f denotes the tensor which is constant with respect to axes Oxyz
so that
*0
91 = 0, (9.6.8)

Furthermore, it is natural to introduce into consideration tensor 9f2 de-


noted for brevity as 9 2 . By virtue of eq. (4.4.2), we have

(9.6.9)

and using the formula for differentiation of tensor (4.3.13) we obtain

Here Wi denotes the angular velocity vector of the carried body referring
to axes Oxyz fixed in the carrying body.
According to (4.8.11) the principal moment K? of the relative momenta
about pole 0 is equal to

(9.6.11)
so that its time derivative with respect to axes Oxyz is given by

(9.6.12)

The relationships (10) and (4.3.12) were taken into account during the
derivation. Using now equalities (9)-(12) we expand the expression needed
9.6 Relative motion of rigid bodies 477

for eq. (1.26)

*0 *0
8 ·W +w X K? + Kr = w' X 82 . W - 8 2 . (w' X w) +

(9.6.13)

By means of eqs. (7) and (9) we have

and
• I *I
W=W +w xw' . (9.6.14)

Now substituting the latter equations into (1.26) and using simple trans-
formations we arrive at the following vectorial equations of rotation for the
system under consideration

8f . w + w X 8f . w + Mlr~l X Wo + 8 2 . (w + W') +
(w + w') x 8 2 . (w + w') + M2r~2 x wC2 = mo. (9.6.15)

Here, for the sake of brevity, the following notation for the vectors of ac-
celeration of the pole 0 and the centre of inertia

Wo =vo +wo x vo,


* } (9.6.16)
. ,
WC2 = Wo + w x rC2 + wx (w x r~J + 2wx r*'C2 + **'
r C2

is introduced. Equation (15) could be obtained by applying the theorem on


the change in angular momentum of the system of two bodies and taking
into account that the vector of the absolute angular velocity of the carried
body is equal to w + w'.
It is also worth noticing the following identity

a x 8· b - b x 8· a = (8 - 19E) . (b x a) , (9.6.17)

which is valid for any symmetric tensor of second rank. The first invariant
of this tensor, i.e. the sum of its diagonal components, is denoted as

(9.6.18)

and E is the unity tensor. Using this identity we find

(W+W') x 8 2 . (W+W') w x 8 2 . W + w' x 8 2 . w' + (9.6.19)


2w' x 8 2 . w+ (8 2 -19 2E) . (w' x w).
478 9. Dynamics of relative motion

This result simplifies the calculation of the term in eq. (15). The matter is
that vector w must be projected on the axes of the carried body in which
vector w' and tensor 8 2 are given. With this in view it is reasonable to
separate the terms of the type w' x 8 2 • w' which are easy to calculate since
all these quantities are given in the axes of the carried body.
We proceed to constructing the equation of the carried body. They are
split into two sets: the first set for coordinates ql, Q2, Q3 describes the relative
motion of the centre of inertia of the carried body and the second set for
coordinates Q4, Q5, Q6 describes the relative rotation.
We can omit the terms which do not depend on the generalised coordi-
nates because expressions (2.9) and (2.10) for the potential energies must
be differentiated with respect to the generalised coordinates to yield the
equations of motion (2.19). Hence, taking into account eqs. (1), (7) and (9)
we have

no = M2 (~o +w x va) .rC2 , }


(9.6.20)
nw = 1
-2M2 1w x r '12
C2 1
- 2W· 8 2· w.

From eqs. (11) and (4) we obtain for s = 1,2,3

(9.6.21)

and, by eq. (2.6), the first set of the equations of motion for the carried
body take the form

(9.6.22)

Notice that in the case under consideration

1
T, = -M2 *' . rc
rc *' +-w'·
1 8 2 . w' ' (9.6.23)
r 2 2 2 2
so that

Making use of transformation (2.3) we find as expected that

(9.6.24)
9.6 Relative motion of rigid bodies 479

The equations of motion (22) take the form

(9.6.25)

where the vector of absolute acceleration WC2 of the centre of inertia of the
carried body is given by expression (16).
Let us consider now the set of equations of the carried body for the
angular coordinates q4, Q5, Q6· The calculation order is prompted by eq.
(2.19). Using expressions (20) we obtain

(9.6.26)

Expressions for the derivatives of the inertia tensor with respect to the
coordinates are easy to obtain using relationships (10). To this end, we
notice that the angular velocity vector Wi is a linear combination of the
generalised velocities Q4, Q5, Q6

(9.6.27)

with the coefficients

(9.6.28)

depending On the generalised coordinates Q4, Q5, Q6. For example, if Euler's
angles are taken as the generalised coordinates, i.e.

then vector Wi is represented in the form


I n·/. 00
w
0
=13o/+nu+13'P,
01'
(9.6.29)

where i 3 , n, i~ are the unit vectors of axis Oz, the nodal line and axis C2 Z ' ,
respectively. In this case

Applying formula (27) we can cast eq. (10) in the form


480 9. Dynamics of relative motion

and thus

(9.6.30)

The generalised centrifugal force has the form


. 1
Q~ = '2w. (es x 8 2 - 8 2 xes) . w = (w xes)· 82· w. (9.6.31)

Furthermore using expression (11) for vector K? we obtain byeq. (2.14)

Qw = -woaK? ° 8
s oqs- = -w· 2· e s·
.- (9.6.32)

Determination of the generalised gyroscopic forces reduces to calculating


the Eulerian operator over vector K?, i.e. turning to eq. (11) we have

Thus we need e* s . Applying formulae for differentiating the unit vector of


the coordinate axes of the carried body

(9.6.34)

we take the derivative of the first equation with respect to ql and that of
the second equation with respect to time, bearing in mind that formulae
(1.3.11) is valid

since vectors i~ depend only on the generalised coordinates. The result is

Accounting for the identity

e s x (w ' x 0' ) -
lk W ' x(e s x lk
0' ) = lk
0' X W (' x e s ) ,
we can recast the above result in the form

( e* s -
ow'
oqs - w, x e s ) x 0'
lk = 0, k = 1,2,3,
9.6 Relative motion of rigid bodies 481

yielding

* aw'
es=~+w
I
xes, 8=4,5,6. (
9.6.35 )
uqs
Using formulae (10), (30) and (17) we can transform expression (33) as
follows

E; (K?) = 82 . (Wi xes) + (Wi X 8 2 - 8 2 X Wi) . e s -


(e s x 8 2 - 8 2 xes) . Wi = (28 2 - E'!9 2) . (e s x Wi). (9.6.36)

In addition the generalised gyroscopic forces are

r s = 2w . ( 8 2 - ~ E'!92) . (Wi X es ) (8 = 4, 5,6) . (9.6.37)

The gyroscopic coefficients "Isk are, due to eqs. (2.17) and (27),

"Isk = 2w· (82 - ~E'!92) . (ek xes) (k,8 = 4,5,6). (9.6.38)

We arrive at the following differential equations of rotation of the carried


body
Es (Tr) = Qs + (w xes)· 8 2 . w - w· 8 2 • e s +
2w· ( 8 2 - ~E'!92) . (Wi xes) (8 = 4,5,6) . (9.6.39)

In order to expand the right hand side of this equality we take eq. (23) and
apply equalities (10), (30) and (35), to get

Es (Tr ) = (e s · 8 2 • w ')* - -aw' . 8 2 . WI - -w a8-2 . w I


1 I .-
aqs 2 aqs
I
= e s .8 2. W +e s · (Wi X 8 2 - 8 2 X Wi) . Wi + (Wi xes) .8 2 . Wi -
(*
1 I . ( e s x 8 2 - 8 2 X e s) . wI = e s · 8 2. w I +w I x 8 2 . W' ) . (9.6.40)
"2w

Expression (39) can now be recast in the form

es . [8 2. Wi +w' x 8 2 . Wi + w x 8 2 . W + 8 2 . w+

2w' X (82 - ~E'!92) . w] = Qs (9.6.41)

w
Replacing here by Wi according to formula (14) and using identity (19)
we arrive at another form of equations for the rotation of the carried body
e s · [82 . (W+W') +(w +W') x 8 2 . (w +W')] = Qs (8 = 4,5,6)
(9.6.42)
482 9. Dynamics of relative motion

The value in brackets is the left hand side of the vectorial equation for
rotation of the carried body under absolute motion

(9.6.43)

the right hand side denoting the principal moment of the external forces
acting on the carried body about its centre of inertia C2'
Varying the generalised coordinates q4, Q5, Q6 implies that the carried
body is subjected to an infinitesimal rotation relative to the carrying body
which is described, due to eq. (27), by the following vector
6
()' = L e s8Qs. (9.6.44)
s=4

Referring to eq. (5.2.5) we can conclude that


6
L Q s8qs = m C2 . ()', Qs = m C2 . e s (8 = 4,5,6). (9.6.45)
s=4

Equation (42) is easy to obtain from the vectorial equation of rotation


(43). However equations (39) and (41) are written in a way that displays
,
more clearly the physical meaning of their separate terms. Moreover, the
derivation is simpler, for example, there is no need to calculate ~ , only w'.
The equations of motion of the carried body (25) and (42) in Lagrange's
form can be easily written also in the form of the Euler-Lagrange equations.
Let us take the projections of vectors r*'C2 and w' on axes C2 x'y'z' as the
quasi-velocities. The time-derivative of the vector with respect to the axes
of the carried body are denoted by a cross over the letter, then

r*'C2 =r+'C2 +w, x rc


, , w*' =W
2
+' +w, x w ,+'
=W (9.6.46)

The generalised forces on the right hand sides of eq. (25) should now be
replaced by projections of the resultant vector of external forces applied to
the carried body on axes C 2 x' y' z' fixed in the body. Then we obtain three
equations of motion in projections on these axes which can be cast in the
form of the single vectorial equation

M2 [+ ' *
Vr +w x v r + Vo +w x Vo + w.x,rC2 +

w x (w x r~J + 2w x vrl = V2, (9.6.47)

where Vr =rc*' 2 is the vector of the relative velocity of point C2 .


In the equations of rotation we should replace Qs and e s by projections
of the resultant moment m C2 on axes of the carried body and the unit
9.7 Examples 483

vectors i~, respectively. Performing these replacements in scalar equations


(41) we write down the result as the single vectorial equation of rotation
of the carried body

8 2 . (~ + ~') + w X 82 . W + w' X 82 . W' +


2w' x ( 8 2 - ~E'l92) .
W = m C2 . (9.6.48)

9.7 Examples
9.7.1 Equations of rotation of thf rigid body with a fixed point
accounting for the rotation of the Earth
Let us view the earth as the carrying body and introduce the basis axes
Oxyz using the unit vectors el, e2, e3 of the geocentric coordinate system
described in Sec. 9.5. The origin 0, which is a point on the surface of
the earth, is taken as the pole of the axes Ox' y' z' fixed in the body. The
equations of motion (6.28) are derived under the assumption that the pole
coincides with the centre of inertia of the carried body. They are also valid
if the pole is immovable (v r = 0) which is the case in question.
The axes Ox' y' z' are assumed to coincide with the principal axes of
inertia at pole 0, that is
(9.7.1)
The angular velocity vector w of the carrying body has both constant value
and direction. One should set ~ in eq. (6.48). The projections of won the
axes of the carrying body and the axes of the geocentric system are denoted
by Wsand W s, respectively. Thus
3
Ws = W . i~ = L Wkek . i~, (9.7.2)
k=l

where
WI = U cos <P, W2 = 0, W3 = U sin <P, (9.7.3)
<P and U denoting respectively the latitude and the angular velocity of the
earth.
The term in eq. (6.48) due to the Coriolis acceleration is

w' x [(A - B - C) i~i~ + (B - C - A) i;i; + (C - A - B) i~i~]. w


= [W;W3 (C - A - B) - W~W2 (B - C - A)]i~ + [w~wdA - B - C)-
W~W3 (C - A - B)] i; + [W~W2 (B - C - A) - W;WI (A - B - C)] i~.
(9.7.4)
484 9. Dynamics of relative motion

The other terms are as in Euler's equations. We arrive at three equations,


the first having the form

Aw~ + (C - B) (w;w; + W2W3) + A(W2W; - W3W;) +


(C - B) (W;W3 + W;W2) = ml, (9.7.5)
and the others being obtained by a cyclic permutation of the letters and
indices.
The terms of the type W2W3 originating from the centripetal forces of
inertia are proportional to U 2 and can be cancelled out. Their influence
can be taken into account by keeping the corresponding terms in expression
(5.6) for the potential energy of the gravity force.
Finally we arrive at the following equations

Aw~ + (C - B) (w~w~ + W~W3 + W~W2) + A(W2W~ - W3W~) = ml, }

Bw; + (A - C) (w~w~ + W~Wl + W~W3) + B (W3W~ - WlW~) = m2,


Cw~ + (B - A) (w~w~ + W~W2 + w~wd + C (WlW~ - w2wD = m3.
(9.7.6)

9.7.2 Heavy top


As an example of the application ofthe general equations (6) let us consider
the influence of the rotating earth on the motion of a heavy top. Let us
determine the position of its axis of symmetry 0 Zl by angles 0: and (3
described in example 7.18.2 and apply the system of axes n, n/, i~ "half-
bounded" to the heavy top. As usual, <p denotes the angle of rotation about
axis OZ'.
Equations (6) are written under the assumption that axes Ox'y' Zl are
fixed in the body and, for this reason, the second equality in eq. (6.46) was
used for the derivation. Now this equality should be replaced by

W
*, =W+' +Axw' , (9.7.7)
I
where (:t; denotes the vector whose projections on axes n, n/, i~ are equal to
the time derivatives of the projections of vector w' on these axes (denoted
by w~ as above). The vector of the angular velocity of the "half-bounded"
axes A differs from w' in the absence of the spin component, i.e.

(9.7.8)
Now
*, +' °1· I +'
W =W -1 3 <P X w =W -<pn Wl
. I I
+ <pnw2
• I

and the quantities w~,w;,w; in eq. (6) should be replaced by w~ + cpw;,


w; + cpw~, w;.
9.7 Examples 485

The role of axes f" 'f/, ( in the table of direction cosines (7.18.20) are
the northern and western directions and the upward vertical, respectively.
Then, according to eqs. (2) and (3), we have

wn WI = U ( - sin <I> sin (3 + cos <I> cos (3) = U cos (<I> + (3) ,
wn ' W2 = -U sin (<I> + (3) sin a, Wi~ = W3 = U sin (<I> + (3) cos a.
The moment of the weight of the heavy top about the point of support
is

rnO -i;z' x e3Q = Qz' (-n sin (3 - n' sin a cos (3 + i; cos a cos (3) x i;
Qz' (n' sin (3 - nsinacos(3).
Accounting that A = B, differential equations (6) become

A (w~ + cpw~ + W2W~ - W3W~) + (0 -


+ W~W3 + W~W2) )
A) (w~w~
-Qz' sin a cos (3,
=
A (w~ - cpw~ + W3W~ - WIW~) - (0 - A) (w~w; + W;Wl + W~W3)
= Qz' cos (3,
o (w~ + WIW~ - w2wD = o.
(9.7.9)

In the case of a nearly vertical top, the angles a and (3 as well as the
projections w~, w~ of the angular velocity vector are small. Also Ws being
proportional to U are small. Neglecting the products of small quantities,
we reduce the first two equations (9) to the form

A (w~ + cpw~ + W2W~) + (0 - A)(w~ + W2)W~ = -Qz'a, }


(9.7.10)
A (w~ - cpw~ - WIW~) - (0 - A) (w~ + WI) W~ = Qz' (3,

and, by virtue of the third equation,

W~ = const.
Within the adopted accuracy of calculation we are to accept that

WI = U cos <1>, W2 =0, W3 = U sin <1>,


W~ = -n, , .
W2, -- (3. , w3 = 'P.
The system of differential equations (10) can be now written in the form
of a single linear second order differential equation

A (.8 + iii:) - Oicp (.8 + in) - Qz' ((3 + ia) = OcpU cos <1>. (9.7.11)

It has a particular solution

(9.7.12)
486 9. Dynamics of relative motion

corresponding to deviation of the axis of the top to the south (if (p > 0 and
Zl > 0). This motion is imposed by oscillations with frequencies

(9.7.13)

Positiveness of the radicand ensures that these frequencies are real-valued


and is the necessary condition for stable rotation of the top. It reduces to
the inequality

C(p> 2JQA z l. (9.7.14)

The more general condition accounting for the masses of the suspension
rings is given by formula (7.9.24).

9.7.3 Rigid body carrying rotating flywheels


The carrying body is assumed to have a fixed point 0. Let the axes Oxyz
be bound to the carrying body. The flywheel rotates with angular velocity
w' = (pi~ about an axis which is fixed in the carrying body. The centre
of inertia of the flywheel lies on this axis and its position referred to the
carrying body is prescribed by the radius vector rc.
In eq. (6.6) we have now

Vo = 0, WC2 = Wc = wx ( w x rc + w x rc.
'). I

The term M2rc x Wc is transformed as follows

wx M 2 (Erc·rc-rcrc
I I I ')
·w+ M 2 (E rc·rc-rcrc
I I I ')
·w.

Let eO denote the tensor of inertia of the system consisting of the carrying
body and the "frozen" flywheel at point 0, then

Using eq. (6.15) we can recast eq. (6.15) in the form

eO . w+ w x eO . w + (p2i~ x ef . i~ + 2(pi~ x ef . w+
(p (ef - E't9f) . (i~ x w) + ef . ((pi~t = mO. (9.7.15)

Since the flywheel is a body of revolution we denote its central moments


of inertia by e 1= e 2, e 3, to get
9.7 Examples 487

Thus
., e C2.13.,
13 X 0,
e c2 . ('1'13
.. ')- ef . (0i~ + ifiw X i~) = e30i~ + e 1 ifiw X i~

and

2ifii~ ef . w + ifi (ef - E'!?f) . (i~ X W) + ef . (ifii~)-


X

2e1ifii~ x w + ifi [e 1 - (2e 1 + e 3)l i~ x w + e30i~ + e 1ifiw x i~


e30i~ + e 3 ifiw x i~. (9.7.16)

Equation (15) is rewritten as follows

ea. w+ w x eO . w + e 3 (0i~ + ifiw x i~) = rna. (9.7.17)

Let us assume the coordinate axes to be coincident with the principal axes
of tensor eO, then

Then projecting eq. (17) on these axes we arrive at three differential equa-
tions

AWl + (C - B) W2W3 + e 3 [a0 + (W2')' - w3(3) ifil = me;, }


BW2 + (A - C) W3Wl + e 3 [f30 + (W3 a - WI ')') ifil = m~, (9.7.18)
CW3 + (B - A) W1W2 + e 3 h0 + (W1f3 - W2 a ) ifil = me;,

where a, f3, ')' denote the cosines of the angles between direction i~ and axes
Oxyz.
The equation of rotation of the carried body, i.e. the flywheel, is con-
structed in the form of eq. (6.41). Under the adopted notation

aw' .,
ecp = aifi = 13

and furthermore

.,
13· e C2 · w*'
., e C •
13 . -2 . W

The other terms in the brackets in eq. (6.41) form the vector which is
orthogonal to i~ and do not appear in the equation for rotation of the
flywheel. The latter has the form

(9.7.19)
488 9. Dynamics of relative motion

which one might easily expect. Equations (18) and (20) were derived by
Volterra. In the case of n flywheels these equations are as follows
n
AWl + (C - B) W2W3 + L e~ [0kD:k + <Pk (W2f'k - w3,8k)] = me;,
k=l
n
BW2 + (A - C) W3W1 + L e~ [0k,8k + <Pk (W3D:k - Wl"Yk)] = m~,
k=l
n

CW3 + (B - A) W1W2 +L e~ [0k f'k + <Pk (W1,8k - W2D:k)] = m~


k=l
(9.7.20)
and correspondingly

e~ (0k + W1D:k + w2,8k + W3f'k) = Q'Pk (k = 1, ... , n). (9.7.21 )


An expression for the kinetic energy of the carrying body with the fly-
wheel obtained in Subsection 4.12.2 is given by formula (4.12.3). Trans-
forming it according to the present notation we obtain

1 W · eO
T ="2 -1 .W+"21 (W+1./3CP.) . eO
-2· (W+1./3CP.) +"21 M 21wxrc/ 12
1W .
="2 [eO eC M 2(E rc·rc-rcrc
-1 + -2 +
/ / / /)] 1 e ·2 +W·
.W+"2-3CP e-2O·13CP
./ .
1 .e
= "2W ° . W + "2e3CP
1 . 2 + (WID: + w2,8 + w3')' ).cpo

In the case of n flywheels

T = 1
"2w. e ° . W + "21~
~ e3 CPk + 2 (W1D:k + w2,8k + W3f'k ).]
k[.2 CPk· (9.7.22 )
k=l
It is natural to take the moments of the active forces about axes Oxyz
in eq. (20) as independent of the rotation angles CPk of the flywheels.
The generalised forces Q'Pk are assumed to be dependent on the quantities
characterising the position and motion of the carrying body. This can be
realised by means of the active systems measuring the above quantities and
producing the moments applied to the flywheels. The coordinates CPk are
then quasi-cyclic and the corresponding momenta are

(k=I, ... ,n).


(9.7.23)

The Routhian function, due to eq. (7.9.13), will have the form

(9.7.24)
9.7 Examples 489

and yields the Euler-Lagrange equations


n
AWl + (C - B) W2W3 +L {(W2'Yk - w3f3k) [Pk - e~ (wlak + W2f3k+
k=l
W3'Yk)] + ak [QCPk - e~ (wlak + w2f3 k + W3'Yk)]} = m~. (9.7.25)

The other two equations are obtained by means of a cyclic permutation


of the letters and indices. It is also necessary to add eq. (7.19.2)

Pk = QCPk (k = 1, ... ,n). (9.7.26)

If 'Pk are cyclic coordinates then the generalised forces are absent and
the momenta Pk are constant. The problem is reduced to considering the
system of three equations (25) and three kinematic relationships determin-
ing WI, W2, W3 in terms of the angular coordinates of the carrying body and
the corresponding generalised coordinates.
Returning to the general case we construct the relationship

n
L QCPk (wlak + w2f3k + W3'Yk) = m~wI + m~w2 + m?w3
k=l

which is easily obtained from eq. (25). The first line represents the quadratic
form R2 of the Routhian function, whereas the expression in the second line
can be cast by means of eqs. (26) and (23), as follows
n
L QCPk (wlak + W2f3k + W3'Yk)
k=l

TUrning now to formula (7.9.15) we can obtain the power of the active
forces

(9.7.27)

The example of a body carrying flywheels demonstrates clearly the con-


cept of cyclic coordinates. The carrying body can be represented as a
closed shell, the rotating flywheels are covered but their presence essentially
changes the motion of the carrying body since it results in the appearance
of additional gyroscopic terms in the equations of motion for the positional
(explicit) coordinates.
490 9. Dynamics of relative motion

9.7.4 Oscillations of particles attached to a moving rigid shell


Let N denote the number of carried particles. Taking into account eqs.
(6.14) and (6.19) and considering the centre of inertia C of the shell (car-
rying body) as the pole 0, we can write the differential equation of rotation
(6.15) in the form

(e + t,e
G Gi ) ·w+wx (e + t,eG Gi ) ·W+

" [e ~
i=l
N
Ci . :;/~ +w''t x e + e Gi . w' 1,
2w' x
1,
Gi . w-

(9.7.28)

-=-=>
Here r~ = CCi stands for the position vector of the centre of inertia C i of
the i - th carried body. The position vector ra of the joint centre of inertia
of the shell and the carried bodies is determined by the equality

(9.7.29)

and the absolute acceleration of point G is equal to

WG WG +Wx r~ + wx (w x ra) +
1
L mi (2W x ;: + *;:) .
N
---N:-:-- (9.7.30)
M + Lmi i=l
i=l

Let us express m G on the right hand side of eq. (28) in terms of the
resultant moment m G about the centre of inertia G

mG mG + r~ x V = mG + (M + tk=l mk) ra x WG

N
mG + Lmkr~ x [WG +w x r~ +wx (w x r~)] +
k=l
N

r~ x Lmi (2WX;: + '1-*:). (9.7.31)


i=l
9.7 Examples 491

Inserting the latter equation into eq. (28) and replacing Wi by its expression
we obtain

(e C + ~eCi) .w+wx (e c + ~eCi) ·W+ ~ [2w~ x eCi·w


- (efi + ef2 + ef~ ) w~ x w + e Ci . w: +w~ x e Ci . w~] +
N
L mi r i x {w x (r~ - r;;) + w x [w x (r~ - r;;)]} +
i=l
N

"L-- mi (r~ - (*' **') = m


r;;) x 2wx r i + r i G. (9.7.32)
i=l

Let us consider the simplest case in which a single carried body oscillates
along a straight line within the shell of a satellite, [76). The equation of
oscillation of the centre of inertia C 1 of this body is given by

r' = rb + aec (t) , (9.7.33)

where the mean value offunction c (t), within a sufficiently long time inter-
val, is zero. Then rb determines the mid-position Cp of the centre of inertia
of the body relative to the shell. The unit vector prescribing the direction
of the above straight line is denoted by e and the oscillation amplitude is
denoted by a. We can take m G = 0 by assuming the resultant vector of
the gravity force to be applied to the centre of inertia G and neglecting the
resisting forces of the atmosphere.
We have then that
M
r;; = M r' - r = m r' w' = 0,
+m r',
(9.7.34)
G M+m'
and the equation of motion (32) can be written in the form

(eC+ eCl ) . w+ w x (eC+ eCl ) . w


: m r' x [w x r' + w x (w x r') + 2aEw x e + aeE:). (9.7.35)
+m
We replace here r' by its expression (33), put the term

Mm
- M rb x [w x rb
+m
+ wx (w x rb))
onto the right hand side and introduce the tensor

eGo + eGo
1
= eC+ eCl + MM+m
m (Er'.
0
r'0 - r'0 r'0 ) , (9.7.36)
492 9. Dynamics of relative motion

which is equal to the inertia tensor of the shell and the body" frozen" III
the shell at the joint centre of inertia Go. Indeed, by eq. (4.4.2)

eGo e C + M (Er~o . r~o - r~or~o) ,


e{O e Cl + m [E (r~-r~J . (r~-r~o) - (r~-r~o) (r~-r~J] ,
and we arrive at eq. (36) by adding these equations and using formulae
(34).
Equation (35) can be rewritten as follows

( eGo + e{o) . w+ w x (eGo + e?o) . w = L Go (9.7.37)

Vector LGo can be referred to as the perturbing torque about point Go


caused by the motion of mass m. Its expression reduces to the form

LG 0 = - M2M
+mmaE: (t ) { [ , .e
Ero -"21 ( roe + ero
I I )]
. w+

wx [Er~.e-~(r~e+er~)] .w}- :+mma2E:2(t)[(E-ee).W-

W x ee . w ] - 2M m aE:. (t ) ( ' e - eroI ) . w-


Ero·
M+m
~Mm a2c(t)E(t)(E-ee).w- Mm aE(t)r~xe. (9.7.38)
+m M+m
In this expression only the terms with non-zero mean value are essential.
They are

L*Go Mm 2 -
= __- - a E: 2 [(E - ee)· W - w x ee· w]. (9.7.39)
M+m
It is assumed that the mean value is calculated over the time interval
containing a sufficiently large number of periods of E: (t) . However this time
interval should be sufficiently short so that one can neglect the changes in
wand w within it.

9.8 Equations of motion of a rigid body having a


cavity filled by fluid
This problem, [98], is studied here as an example of the application of the
equations of the dynamics of relative motion. The carried body is a fluid
whose motion relative to the carrying body is prescribed.
Our attention will be restricted to the case of single connected cavity in
a rigid body. Let us recall that the domain is said to be single connected if
9.8 Equations of motion of a rigid body having a cavity filled by fluid 493

any closed curve in it can be reduced to a point by means of a continuous


transformation. For instance, the domain in the sphere is single connected
whereas that in a torus is double connected. The fluid, that fills the cavity,
is assumed to be ideal, incompressible and homogeneous. Then its flow is
irrotational and we can introduce the velocity potential <1>1 (e, 'fJ, () which
is a harmonic function in the coordinate basis Oe'fJ(. The gradient of this
potential yields the vector of the absolute velocity Va of a fluid particle.
Let Oxyz be the axes fixed in the body, then using equations for trans-
formation of the coordinate systems we can express the velocity potential
<1>1 in terms of x, y, z. Clearly, time will appear explicitly in this dependence

<1>1 (e, 'fJ, () = <I> (x, y, z, t) .


Thus,

Va = grad <1>, (9.8.1)


<I> remaining a harmonic function of x, y, z.
The normal components of velocity of the cavity wall S and the fluid
particles on S coincide. Denoting the unit vector of the inward normal to
S by n we have
Va· n = n· grad <I> = (vo + w x r') . n = Vo· n + w· (r' x n), (9.8.2)
since the geometric sum Vo +w x r' is the velocity vector of the point ofthe
rigid body on Sand r' is the position vector of the cavity surface. Notice
that

grad (vo . r') = Vo, (9.8.3)


as vector Vo does not depend on the coordinates x, y, z.
We enter the harmonic vector B (x, y, z) that is the vector whose projec-
tions B 1, B 2, B3 on axes Oxyz are harmonic functions. They are determined
within the cavity V by means of their normal derivatives on surface S

(9.8.4)

Here n1, n2, n3 denote the projections of n on axes Oxyz and


oBg oBg oBs oBg
-- = --n1 + --n2 + --n3 (
s = 1,2,3)
on ox oy OZ
denote the normal derivatives of functions Bg.
It is known that Neumann's problem, i.e. determination of harmonic
function prescribed by its normal derivatives on the boundary, has a unique
solution when the following condition

J oBsdo
on
= 0 (9.8.5)
494 9. Dynamics of relative motion

holds. Here do denotes the surface element and integration is over the
surface 8 bounding the domain. This condition is met. In order to prove
this, it is sufficient to remember the formula for transforming the surface
integral into a volume integral

(9.8.6)
where dT denotes an element of the volume V and cp is a continuous function
having continuous partial derivatives of first order in the volume V and on
the border 8. In our case

JoBsan J
s
--do =
s
(yn3 - zn2) do = J(OYoz
v
- - -OZ)
oy dT == 0,
which completes the proof.
Conditions (4) can also be cast in the form

n . grad B = r' x n. (9.8.7)


Here we introduce tensor grad B which is dyadic product of the operator

and vector B. Notice that in particular grad r' = E, E being the unit
tensor.
Given vector B, the velocity potential <I> can be found by the relationship

<I> = Vo . r' +B . w. (9.8.8)


It is a harmonic function of x, y, z as r' and B are harmonic vectors whereas
Vo and ware independent of the coordinates. Time appears in <I> only
by means of these coordinates. This means that if the vessel with a fluid
initially at rest begins to moves and then stops, then the containing fluid
stops at the same time instant.
Function <I> satisfies the boundary condition (2). Indeed,

va = grad <I> = grad (vo . r') + grad (B· w) = Vo + grad (B· w), (9.8.9)

and then, using eqs. (3) and (7), we find

n· Va = n· grad <I> = n· Vo + (r' x n) . w, (9.8.10)


which is required.
9.8 Equations of motion of a rigid body having a cavity filled by fluid 495

Denoting the velocity of the fluid particle relative to the vessel, i.e. axes
Oxyz, by r*' we have

Va = Vo +W X r ,+ r*' ,
and comparing with eq. (9) yields

r*' = grad (B . w) - W x r'. (9.8.11)


Therefore, if the motion of the vessel is prescribed we also know the
motion of the fluid which fills the vessel completely. It suffices to have the
solutions of three Neumann's problems which depend only on the form of
the vessel.
We proceed now to construction of the equations of motion of the system,
that is the rigid body (vessel) and the fluid filling it. These equations are
(1.17) and (1.26). The first equation contains the relative velocity rc*' and
**'
the relative acceleration r c of the centre of inertia of the system. However,
the vessel is filled completely and the fluid is homogeneous, i.e. its motion
within the vessel does not affect the position of the centre of inertia of the
vessel
*' 0,
rc= **'
rc= 0,
and thus eq. (17) does not differ from the equation of motion of the rigid
body
M [vo +W x Vo + wx (w x r~) +W x r~] = V, (9.8.12)
with M denoting the mass of the system.
*0
The tensor of inertia e of the system at point 0 remains constant in
eq. (1.26) for the same reason. It coincides with that of the system with
the "frozen" fluid. Thus
*0
e =0
and all we seek is the moment of the relative momenta K? and its time-
derivative. By virtue of eq. (11) we have

K?=p jr'x;' dT=P jr'xgradB.WdT-P jr'x(wxr')dT.


v v v
(9.8.13)
Here pdT = dm denotes an element of the fluid mass and p its mass density.
Furthermore

p j r' x (w x r') dT = W . j (Er' . r' - r'r') dm = W . e~ (9.8.14)


v v
496 9. Dynamics of relative motion

is the angular momentum of the" frozen" fluid about pole 0 and e~ de-
notes the inertia tensor at point o.
Now we proceed to calculation of the integral

GO =p J
v
(r'x gradB· w) dT = K? + e?· w. (9.8.15)

As follows from eq. (4.8.8), vector GO is the angular momentum of the


fluid in the vessel provided that point 0 is fixed.
Denoting the tensor of inertia of the rigid body (without fluid) at point
o by

and noticing that

we can write eq. (1.26) in the form

ef· w + w x ef· w + Mr'c x * +w x Vo )


( Vo + G* ° +w x GO = mO.
(9.8.16)

Let us find the projection of vector GO on axis x. In principle, we could


operate with vectors but this would complicate the derivation. We have

(9.8.17)

The integrals are easily transformed by means of formulae (6) and (4) to
the following form

Then introducing the notation

(i, k = 1,2,3) , (9.8.19)

instead of (17) we obtain

G~ = Q~I WI + Q~2W2 + Q~3W3 (9.8.20)

and two analogous expressions for G~ and G~.·


9.8 Equations of motion of a rigid body having a cavity filled by fluid 497

Applying transformation (6) and taking into account that Bi are har-
monic functions we obtain

This also means that Qfl = Q~i.


Projections of vector GO are linear functions of projections of vector w.
Thus quantities Qfl are the components of the tensor of second rank. This
symmetric tensor of second rank can be termed as the tensor of inertia of
a rigid body which is equivalent to the fluid in the cavity at point 0 since
the angular momentum of the fluid about point 0 can be represented in
the form

(9.8.22)

similar to the angular momentum of the rigid body having the tensor of
inertia QO at point o.
Equation of motion (16) is now cast as follows

(ef + QO) . w+w x (ef + QO) .w + Mr~ x (~o +w x v o) = rn o ,


(9.8.23)

whilst in the case in which the fluid was "frozen" it would have the form

(ef + e~) . w+ w x (ef + e~) . w + Mr~ x (~o +w x vo) = mO.


(9.8.24)

As an example let us consider the case of a vessel having an elliptic cavity.


The equation of the bounding surface is

It is known that
1 8F 2x 1 8F 2y 1 8F 2z
nl = ~ 8x = a2~' n2 = ~ 8y = b2~' n3 = ~ 8z = c2~'

where
498 9. Dynamics of relative motion

The first of these boundary conditions is written as follows

OBI X OBI Y OBI Z ( 1 1)


ox a2 + oy b2 + OZ C2 = yz c2 - b2 '

and is satisfied by the following harmonic function

b2 - c2
BI = -2 b 2YZ.
+c
By analogy, we obtain

Calculation by eqs. (21) yields

Here A, B, C denote the moments of inertia of the fluid in the vessel

For a sphere a = b = c, and the harmonic vector B is identically equal


to zero. This implies that the ideal fluid is not involved in rotation by the
rotating spherical vessel.
The case of a rotating cylindrical vessel was considered by Zhukovsky in
[98].

9.9 Equations of motion for a solid


In this section the carried bodies are the particles of an elastic solid vi-
brating about the positions which they would possess in a rigid body. This
"rigid skeleton" forms the carrying body. The axes Oxyz whose motion is
described by the velocity Vo of the pole and the angular velocity vector w
are bound to this carrying body. The position of point M of the skeleton
is given by its radius vector oM = p and ro = 00
denotes the radius
vector of the pole 0 relative to the fixed axes with the origin at point ().
The displacement of the particle, with the initial position at point M, is
given by vector u depending upon the coordinate and time. It is essential
that time appears in u only in terms of parameters qQ playing the part
9.9 Equations of motion for a solid 499

of the generalised coordinates. The number of these parameters is taken


as finite and is denoted as n. Remaining in the framework of the methods
of analytical mechanics of systems with finite number of degrees of free-
dom we should adopt that the functional dependence of vector u both on
the coordinates x, y, z of the point in the natural state and the generalised
coordinates

U=U(X,y,Z,q1, ... ,qn) (9.9.1)


is given. This expression can be represented as a power series in terms of
qa
n 1 n n
U = Lqa ua (x,y,z) +"2 LLqaq{3U a{3 (x,y,z) + ... (9.9.2)
0.=1 a=1{3=1
Here we took into account that u =0 at qa = O.
Though only small motions are considered and only the terms linear
in qa are kept in the equations of motion it would be wrong to omit the
quadratic terms just in the expression for u. Some linear terms in the
equations of motion describing certain essential effects would be lost in
this case. However, there is no need to keep the terms of order higher than
the second in eq. (2).
Let us explain the way function u was chosen using the example of an
inextensible elastic rod. We direct axis Ox along the rod axis, then the
bending vibration occurs in planes xy and zx. Under bending the linear
part of the displacement vector of the points of the rod axis is given by
n n

Uo= L i 2 qara (x) +L i3qn+a'l/'a (x), (9.9.3)


0.=1 0.=1
where the vibration modes in planes xy and zx can be taken as ra (x) and
'l/' a
(x), respectively, and qa and qn+a denote two systems of the generalised
coordinates. In order to take into account the rotation of the cross-section
under bending, we denote the angles of rotation of the cross-section about
axes y and z by 12 and 13' then
n n
i212 = -i2 L qn+a'l/'~ (x), i3'Y3 = i3 L q~r~ (x).
0.=1 0.=1
The additional component of vector u, which determines the displacement
of the point of cross-section with abscissa x and the radius vector i 2 y + hz,
is equal to

(i212 + i3'Y3) X (i 2 y + i3 Z)
-h [y~qar~(x)+z~qn+a'l/'~(x)l (9.9.4)
500 9. Dynamics of relative motion

Thus Uo does not exhaust the linear components of vector u. The terms
of the second power are determined from the known formula for the axial
displacement of the beam

2]1/2
~= (~~) + (~:)
x [ 2
x- [ 1+ dx,

where ~,T/,' denote the projections of the displacement of the rod axis
on axes Ox, Oy, Oz, respectively. It is also assumed that ~ = 0 at x = o.
Making use of eq. (3) we obtain

~ ~ -~]o {[t q~cp~ (X)]


a-I
2
+ [t qn+a'l/J~
a-I
(x)]2} dx.

Therefore, accounting for the terms quadratic in the generalised coordinates


we obtain the following expression for vector u
n
U = L {qa [-ilYCP~ (x) + hCPa (x)] + qn+a [-hz'l/J~ (x) + ia'l/Ja (x)]} -

1 1
a=1

~;, t, t, [q.qp ,,~(<) "p ('1 ~ + qn+.qn+' ,p~ ('1 "'p ('1 ~].
(9.9.5)
Any form of deflection of the rod axis is known to be represented as
a series (2) in terms of the vibration modes. Looking for an approximate
solution we can take a finite number of the vibration modes, i.e. the terms
in the series. Moreover, instead of the vibration modes it is admissible to
take other functions of x which reasonably approximate the character of the
elastic line of the rod axis. The two above assumptions are justified in prac-
tical calculations of rods and plates on fixed supports. There is no reason
to view these assumptions as unacceptable for construction of the general
equations of motion of solids. The first of these assumptions reducing the
problem to a system with a finite number of degrees of freedom excluding
from consideration the high frequency modes, which are known to be very
difficult to deal with. As is shown below, the second assumption does not
affect the result considerably since the taken functions determine some in-
tegral characteristics which are not sensitive to these functions provided
that their choice is reasonable.
In what follows the integration includes not only distributed parameters
but also lumped masses, point forces etc.
Under the introduced notation the expression for the position vector r'
of the point M with respect to axes Oxyz is
r' = p + u = XiI + yi2 + zia + u. (9.9.6)
9.9 Equations of motion for a solid 501

The position vector of the centre of inertia C is given by the equality

Mr~=Mpc+ J udm=Mpc+ a (q1, ... ,qn), (9.9.7)

so that

(9.9.8)

The equations of motion of the solid are obtained by means of the general
equations of the theory of relative motion (1.17) and (1.26) for the carry-
ing body and eq. (2.19) for coordinates qa. These equations must be linear
is qa and <ia. Thus it is necessary to retain the quadratic terms in those
quantities which should be differentiated with respect to these variables.
These quantities are the kinetic energy of the relative motion Tr , the po-
tential energy no of the inertial forces in a pure translational motion, and
the potential energy nw of the centrifugal forces. Moreover, it is necessary
to take into account the terms linear in qa in the expression for variation
Dr' which is needed for deriving equations for the generalised forces. Thus
we need to keep the quadratic terms in the equation for r'. Account of
quadratic terms in the expression for r' is also required for calculation of
the generalised rotational forces of inertia Q~.
By virtue of eq. (8)

2Tr = J
*' r*' dm
r· = J u* . u* dm =~
~ ~
~ <ia<i(3
0.=1 (3=1
Jau
a
qa
au
. adm,
q(3

and it is clear that for calculation of T it suffices to keep only linear terms
in eq. (2). Then we obtain

Tr ~ t tA
=
0.=1 (3=1
a(3<ia<i(3, A J
a (3 = Va. V(3dm. (9.9.9)

Turning now to eqs. (2.5) and (8) and cancelling out the terms which are
independent of qa we have

(9.9.10)

Here the notation


502 9. Dynamics of relative motion

is introduced. Thus

QOI.o = - 811
0
8qOl. = - (*vo +w x vo ) . (9.9.12)

and neglecting the quadratic terms in formula (2) would lead to a loss of
the linear terms in the expression for Q~.
The tensor of inertia eO which determines I1 W by means of eq. (2.6)
should be found with the same accuracy. Due to eq. (4.3.20), we have

eO J(Er' . r' - r'r') dm

J [E (p + u) . (p + u) - (p + u) (p + u)] dm

J + J
(Ep· p - pp) dm 2 [Ep. u-~ (pu + up)] dm +
J (Eu . u - uu) dm,

and, in order to keep all terms which are quadratic in qOl., we should take
into account such terms in the expression for u. Then we obtain
n n n
eO = e~ + 2L AOI.qOl. +L L QOI.f3qOl. qf3 . (9.9.13)
01.=1 0I.=1f3=1

Here e~ is the tensor of inertia of the rigid skeleton at point 0 and

AOI. J [Ep. UOI. - ~ (PUOI.+UOI. p )] dm,


Qa f3 J [EU a . Uf3 - ~ (U a Uf3 + Uf3U a ) +

Ep. U a f3 - ~ (pU a f3 +U a f3 p) ] dm (9.9.14)

denote the inertia tensors at this point.


Let us proceed to the generalised forces. The prescribed forces, except for
the elastic forces which are discussed later on, are assumed to be dependent
on the coordinates and velocities of the body particles. Thus, accounting
for the changes due to elastic displacements of the body particles we should
adopt the following expressions for the forces
n
F = F* + L (fOl.qOl. + f'0I.4a). (9.9.15)
a=1
9.9 Equations of motion for a solid 503

By F we imply the distributed loads and point forces, F * denotes the


value of F if the solid were a rigid body, and fa and f'a denote the vectors
which are specified in any particular problem and depend on x, y, z. For
instance, the surface aerodynamic forces acting on the airfoil are known to
depend essentially on its oscillation and are expressed in terms of the values
defining its elastic deflection and velocity. The change of forces of hydro-
static pressure on the body immersed in fluid is caused by a change in the
directions of the normal vector to the body surface due to the deformation.
The virtual displacement of the body particle is given by the equality
br bra + br' bra + () x r' + b*r (9.9.16)

t
=

8ro + 8x (p+ qaV") + t, (V" + t, V'"q,) 8q"


and the elementary work is as follows

b'W = JF· brdT = J F* . [bra + () x (p+ ~ Uaqa) +


~bqa (u + ~u~aq~) 1dT+
t
a

J (f/3 q/3 + f,/3q/3) . (bra + () x p + ~ Uaqa) dT. (9.9.17)

The terms quadratic in qa


and qa have been omitted here and dT denotes
an element of the volume, surface or arc.
Let the resultant vector and the resultant moment about pole 0 of forces
F and F* be denoted by V,rno and V*,rn~, respectively. Then the ele-
mentary work is given by
n
b'W = V . bra + rno . () + L Qabqa. (9.9.18)
a=1

t t
Here

V = V *+ (q/3 J f/3 dT + q/3 J f'/3 dT) = V*+ (v/3 q/3 + v'/3 q/3) ,

(9.9.19)

n
rn~ +L (p,/3 q/3 + p,'/3 q/3) . (9.9.20)
/3=1
504 9. Dynamics of relative motion

The constant vectors v{3, v '{3, JL{3, JL'{3 determine the change in the resultant
vector and the resultant moment of the forces due to the body deformation
and are the subject of special investigation, see for example [7].
Expressions for the generalised forces Qo: in eq. (18) take the form

Qo: = J F*· UO:dT +~ [q{3 J


(f{3. UO: + F*· Uo:(3) dT+

q{3 Jfl{3 . UO:dT] = Qoo: + ~ (ro:{3q{3 + r'o:{3q(3). (9.9.21)

The term Qoo: does not vanish when the body is undeformed. This effect is
due to account of the elementary work of "already existing" forces F * due
to the virtual displacement of the body particles from their position in the
rigid skeleton.
Since the generalised coordinates qo: are zero in the natural state, the
expression for the potential energy of elastic forces is a quadratic form of
these variables

(9.9.22)

Taking into account formulae (7) and (8) and notation (11), we arrive at
the following equation of motion of the carrying body (1.17)
n
M [~o +w x Vo +w x Pc +w x (w x pd] = V*+ L (vO:qo: +v'O:qo:)
0:=1
n
-L {qo: [w x aO: + w x (w x aO:)] + 2qo:w x aO: + qo:aO:}. (9.9.23)
0:=1

The case qo: = 0 yields the equation of motion of the centre of inertia of
the rigid skeleton. The additional terms are caused both by changes in the
resultant vector V and the reaction forces acting on the skeleton.
In order to construct the equation of rotation (1.26) of the carrying body
we write down the moment of the relative momenta K? about pole 0

(9.9.24)

where

(9.9.25)
9.9 Equations of motion for a solid 505

The quadratic terms are omitted. By using eq. (1.29) we obtain the moment
of the Coriolis forces of inertia keeping only the terms linear in qa

m~OT = - (e O
·w + w x K~) = -2 ~ qa (w. Aa + ~w x G a ) .

(9.9.26)
It is easy to prove that the above equality is the moment of the Coriolis
forces. Indeed,

w ·Aa + ~w x G a = - Jpx (w x Ua)dm,

where this equation is obtained by means of eqs. (14) and (25) for Aa and
Ga.
Substituting the obtained equalities into eq. (1.26) we obtain

e~ ·w+w x e~ ·w+Mpc x (~o +w x vo)


n n
= m~ + L (I£a qa + I£,a qa ) - L qa [2 (A a . w+ w x A a . w) +
a=l a=l
n n
aa x (~o +w x v o)] - L ria (2w . A a + w x G a ) - L (jaGa. (9.9.27)
a=l a=l
Now we turn our attention to eq. (2.19) for coordinates qa. Expressing the
generalised forces of inertia Q~ of rotation and the generalised gyroscopic
forces r a by means of eq. (2.14), we obtain

cr., ~ -w -f :'>m ~ w J(p+ f,q,U')


r' x x

(ua + t
1'=1
U1'a q1') dm = -w· (G a + t
(3=1
G(3a q(3) , (9.9.28)

where
G(3a = J (U(3 x U a +p x u(3a) dm. (9.9.29)

Expressions for the gyroscopic coefficients "(a(3 are constructed with the
help of eq. (2.18). It is sufficient to retain only the terms independent of
qa as these coefficients are further multiplied by the generalised velocities
q(3. This yields

"( (3
a
= -"((3
a
= 2w· Jor'
-
oqa
or'
x -dm
oq(3
= 2w· J U a x U(3dm = 2w· r a (3
(9.9.30)
506 9. Dynamics of relative motion

FIGURE 9.4.

and equations (2.15) in the form

n n n
LA Q,6 q,6 = L (rQ ,6 q,6 + r'Q,6 q,6 ) - Lc Q,6 q,6 -
,6=1 ,6=1 ,6=1

(a" + t,a",q,) (~o ~ xvo)+w (A" + t,Q""qp )w+


L t,G'"q") + ~ 1, ... ,n).
n
2w . rQ,6 q,6 - w . ( G" + Q"" (0
,6 =1
(9.9.31 )

The notation of eqs. (21), (22) , (11), (14), (30), (25) and (29) is used here
as well as for calculation of the constant vectors and tensors a Q , GQ , AQ
etc. in eq. (31) and in the equations of motion (23) and (27) of the carrying
body. The complexity of the obtained system of equations is a result of the
general statement of the problem. The equations simplify essentially for
particular problems.

9.10 Oscillations of a rotating rod


This problem serves as an example of the application of the general equa-
tions (9.31). Bending vibration of the rotating elastic inextensible rod (a
blade) is considered. The end 0 of the rod is clamped in a turbine wheel
which rotates with a constant angular velocity w about axis OZ1, Fig. 9.4.
The distance between the rotation axis and the clamped wheel is R. The
other end of the rod is free. As the motion of the carrying body (the turbine
wheel) is prescribed, only eq. (9.31) is required. Let us place the pole of
the axes bound to the carrying body to point 0 and direct axis Ox along
9.10 Oscillations of a rotating rod 507

the axis of the undeformed rod. Then

(9.10.1)

and neglecting the weight we can cast eq. (9.31) in the form

n n
L A a,6 q,6 +L (c a ,6 - w 2 Rar,6 - w 2 Q~f) q,6 +
,6=1 ,6=1
n
2w L rg a q,6 = w 2 (A~3 + af R). (9.10.2)
,6=1

It is assumed that the points of the rod axis can move in the tangential
(along axis Oy) and the axial (along axis Oz) directions. In accordance
with eq. (9.5)

ua = -i1Y'P~ (x) + i 2 'Pa (x), u n + a = -i1Z1P~ (x) + i 31Pa (x), )

J'P~ (~) 'P~ (~) d~, u J1P~ (~) 1P~ (~) d~,
x x

U a ,6 = -i1 n + a ,n+,6 = -i1


o 0
(a,{3=l, ... ,n),
(9.10.3)

where

U a ,,6+n = u a + n ,,6 = 0 (a, (3 = 1 , ... ,n ) . (9.10.4)

We have

dm = p (x) dodx,

where p (x) and do denote the mass density of the rod material and the
element of cross-sectional area S (x). Thus the evaluation of the integrals
is reduced to the rule

J J J
I

!(x,y,z)dm= p(x)dx !(x,y,z)do.


o Sex)

As functions 'Pa(x),1Pa(x),p(x),S(x) are taken to be prescribed, the


calculation of all coefficients of eq. (2) is feasible though it can be rather
cumbersome.
The forthcoming analysis will be restricted to the case of a homogeneous
prismatic rod (i.e. p and S are constant). Also, we do not take into account
the terms in eq. (3) originating from the rotation of the cross-section and
adopt that the rod is not twisted. Then the principal axes of inertia of all
508 9. Dynamics of relative motion

cross-sections are parallel to each other. To simplify the analysis we will


consider the case in which these axes are parallel to Oy and Oz, respectively.
Axis Ox is the locus of the centres of inertia of the cross-sections of the
rod.
We take the vibration modes of the cantilever beam rigidly fixed at x = 0
as functions <Po. (x) and '¢o. (x), then under the above conditions

<Po. (x) == '¢o. (x). (9.10.5)


Hence, for the case in question

(9.10.6)

which reduces the volume of calculation by half. Using eq. (9.9), we obtain

Ao.,B = An+o.,n+,B = pSIAo.,B, Ao.,B = J


1

<Po. (s) <P,B (s) ds. (9.10.7)


o
Here and throughout this section the integration is carried out over a non-
dimensional variable s = x/I.
As follows from eqs. (9.11) and (9.14) the projections of ao. on axis ar
Ox as well as the components A~3 of tensor A0. vanish and equations (2)
become homogeneous. This is what we expect since a straight rod on a
rotating wheel can remain straight.
By means of eq. (9.11) we find

J J<p~
1 s

ar,B = a~+o.,n+,B = -pS ds (a) <P~ (a) da. (9.10.8)


o 0

Turning to eq. (9.14) we obtain

Q'if ~ pSI I ["'" (s) "" (s) - s Z1"'" 1


(u) "'_ (u) do- ds,
(9.10.9)
J J<p~
1 s

Q~to.,n+,B = -pSI sds (a) <P~ (a) da.


o 0

One observes a difference between the coefficients of the equations of the


two sets of vibrations which can be explained by the fact that vectors uo.
and un+o. have different directions so that the values
9.10 Oscillations of a rotating rod 509

in coefficients Q~~ (k, s = 1, ... , n) differ from those for k, s = n + 1, ... , 2n.
The potential energy of the bent rod is calculated by means of the well-
known formula

1
IIe = 2Elz JI

v" 2 (x) dx + 2Ely


1 J
I

w"2 dx,
o 0

where v (x) and w (x) denote projections of displacement of the rod axis
on axes Oy and Oz, respectively, and ly and lz stand for the moments of
inertia of the cross-sections about axes Oy and Oz, respectively. In our case

which yields the coefficients

C
a{3 _
-
El
t3
z J "() "()
1

'Pa S 'P{3 s d s -- Elz


t3 c-=et{3 ,cn+a,n+{3 -_ El
t3 Y c-=<JI.{3 . (9.10.11)
o
The mixed coefficients with one subscript less than n and the other greater
than n are absent in eqs. (7)-(11). One can easily see that the gyroscopic
generalised forces vanish in this case. Indeed, turning to formulae (9.30)
and (6) we obtain

(9.10.12)

We enter the notation

(9.10.13)

J(~ + s) J'P~
1 s

m= a {3 m n+a ,n+{3 = ds (a) 'P~ (a) da.


o 0

By means of the following simple formula

JJ JJ
a x a y

dx f(x,y)dy = dy f(x,y)dx,
o 0 0 0
510 9. Dynamics of relative motion

which is valid for 0 ::::: x ::::: a, 0 ::::: y < a, the latter double integral is
reduced to the following single integral

J
1

m a(3 = m n+a ,n+(3 = ~ (1 - u) (2~ + 1 + u) cP~ (u) CP~ (u) du.


o
(9.10.14)

Inserting the obtained results into eq. (2) we arrive at two independent
systems of equations. The first one is for the tangential motion
n
L [Aa(3ij(3 + (>,~ca(3 + w 2 m a(3 + w 2 Aa(3) q(3] = 0 (a = 1, ... ,n)
(3=1
(9.10.15)
whereas the second one is for the axial motion
n
L [A a(3ij(3 + (>,;c a(3 + w 2 m a(3) q(3+n] = 0 ((3 = 1, ... ,n). (9.10.16)
(3=1

As pointed out in Sec. 9.9, it is admissible to take other functions of x,


instead of the vibration modes, which reasonably approximate the character
of the elastic line of the rod axis. Instead of the vibration modes of the
rotating beam we take the vibration modes of the non-rotating cantilever
beam. The latter are determined from the differential equation, e.g. [3],

cP;V (8) - v;CPa (8) = 0 (9.10.17)


subject to the boundary conditions

CPa (0) = cP~ (0) = 0, cP~ (1) = cP~' (1) = O. (9.10.18)


Here

(9.10.19)

>'a denoting the a - th natural frequency and


2 EI
>, = pSl4' (9.10.20)

so that >,2 = >,~ and >,2 = >,; for vibrations in the plane Oxy and Oxz,
respectively. The values of Va are known from the frequency equation and
equal
Jr 3Jr
VI = 1.19372"2' V2 = 0.99610 2 ,
(9.10.21)
2k -1
Vk;::::: --2-Jr, k ~ 4.
9.10 Oscillations of a rotating rod 511

Therefore, the vibration modes of the free axial and tangential vibrations
are identical however the frequencies are different. Their ratio is

(9.10.22)

The vibration modes are known to be orthogonal, that is, the integral
of the products of two different modes over the interval (0,1) is zero. We
normalise the vibration modes such that in eq. (7)

J
1

A:a,6 = CPa (8) CP,6 (8) d8 = 8a ,6, (9.10.23)


o
8a ,6 being the Kronecker delta.
TUrning now to expression (11) we carry out integration by parts

JcP~ JcP~ cp~'


1 1

Ca ,6 (8) CP~ (8) d8 = cP~ (8) CP~ (8) I~ - (8) (8) d8


o 0

I~ + J
1

[cp~ (8) CP~ (8) - CPa (8) cp~' (8)] CP~V


CPa (8) (8) d8,
o
and, referring to eqs. (17), (18) and (23), we find

Ca ,6 - -8a,6V 4a · (9.10.24)

The equations of dynamics (15) and (16) are considerably simplified and
take the form
n
ija+ [(A~)y+W2]qa+W2Lma,6q,6=0 (0:=1, ... ,n) (9.10.25)
,6=1

for the tangential vibration, and the form


n
..
qn+a + ( Aa
\2) z qn+a + W2 'L" m a,6 qn+,6 = 0 (o:=l, ... ,n) (9.10.26)
,6=1

for the axial vibration. Here (Aa)y and (Aa)z denote the eigenfrequencies
of the tangential and axial vibrations for the non-rotating rod, respectively.
The influence of the angular velocity of rotation on the eigenfrequencies and
vibration modes can be revealed. The terms in eqs. (25) and (26) containing
coefficients m a ,6 determine the restoring effect of the centrifugal forces. If
we omit the quadratic term in the expression for the displacement vector
512 9. Dynamics of relative motion

FIGURE 9.5.

u we would not obtain the terms with coefficients m a {3 and the physical
meaning of the problem would be lost to a great extent.
Normalised expressions for functions CPa (x) satisfying eq. (23) are given
by

1
CPa (x) = SIn
. Va COS h Va . h [(sinh Va + sin va) (cosh VaS-
- COS Va SIn Va
cos vaS) - (cosh Va + cos va) (sinh VaS - sin vas)] , (9.10.27)

so that, for a given value of RII , the numerical calculation of coefficients


m a {3 due to eq. (14) presents no problem.

9.11 Equations of motion of a rocket


This problem was studied in [28]. The carrying body includes the perma-
nent structural members of the rocket like shell, equipment etc. along with
the fuel which has not yet burned up to the time instant t under consid-
eration. The carried bodies are the gaseous combustion products in the
nozzle of the rocket. Vector r*'i is the velocity of the i - th particle rela-
tive to the nozzle. The motion of the gas is assumed to be stationary and
straight which allows representation of the vector of the relative velocity r*'
(subscript i is omitted in what follows) of the ensemble of gas particles in
a certain cross-section of the nozzle in the form

r*' = -ev (A). (9.11.1)

Here e stands for the unit vector directed along the nozzle axis to the rocket
head and A denotes the abscissa of the cross-section under consideration,
A = 0 and A = I implying the inlet and outlet cross-sections of the noz-
zle, respectively, Fig. 9.5. The value of the relative velocity v (A) is thus
determined by the nozzle profile and is taken to be independent of time.
In our case, mass M of the rocket decreases due to the exit of gases via
the outlet cross-section and is a prescribed function of time t. The law of
combustion is presumed to be known thus the amount and the configuration
of the fuel are known also. This means that the dependence of the position
vector r~ of the centre of inertia and the inertia tensor eO on time is given.
The mass and the moment of inertia of the gas in the nozzle are neglected.
9.11 Equations of motion of a rocket 513

Let us return to the equations of dynamics of the relative motion. The


sums in eq. (1.17) should be replaced by the following integrals

Mr~ = J *'
r'dm, M rc= J*'r dm, M **'
r c= J**'
r dm. (9.11.2)

In accordance with eq. (1.28) and the definition of vector K? we should


substitute into eq. (1.26) the following

e*° ·W + w x K? = 2 J (
r' x w x r*') dm, Kr
* =° J r' x **'
r dm. (9.11.3)

After these replacements eqs. (1.17) and (1.26) take the form

M[~~+Wxvo+wxr~+wx(wxr~)] +
2w x J*'
r dm + J**'
r dm = V, (9.11.4)

eO ·w+w x eO ·w+Mr~ x (va +w x va) +


2 J (
r' x w x r*') dm + Jr' x **'
r dm = mO. (9.11.5)

As long as we deal with the same aggregate of the bodies this is just another
form of eqs. (1.17) and (1.26). While constructing the equations of motion
for the rocket we should use eqs. (4) and (5) since the derivation of eqs.
(1.17) and (1.26), from the initial equations of motion for a system of
particles, is no longer applicable here. Because of the variable mass, the
second equation in (2) does not follows from the first one and the third
equation does not follow from the second.
We explain the above-said by referring to the derivation of Lagrange's
equations in Sec. 7.1 from the fundamental equation of dynamics. The
following expression

provokes no objection.
The next step was the replacement of the sum by T. In our case T is
determined by formulae (1.5)-(1.7) containing M and eO. However, they
are functions of time now since that part of the fuel which left the rocket
in the gaseous form up to time instant t is not of interest to us. As long as
Lagrange's equations were written in the form (6) it would occur to no one
to differentiate masses mi with respect to time t. However one can forget
it and erroneously take time-derivatives of the integral characteristics like
M and eO relevant to the kinetic energy T.
514 9. Dynamics of relative motion

Turning to eq. (4) we notice that the expression for the elementary mass
dm moving in the nozzle with the relative velocity r*' can be set as follows

dm = -Mdt = v ()...) pSdt = pSd)"', (9.11.7)

where M denotes the flow rate and the mass density p is assumed to be
constant. For this reason, taking into account eq. (1), we obtain

J;' J J
l l

dm = -e v ()...) pSd)'" = eM d)'" = Mle. (9.11.8)


o 0

By virtue of eq. (1) we have


**' .
r = -ev' ()...) ... (9.11.9)
and, furthermore, applying eq. (7)

J J J
l l

*;' dm = Me v' ()...) )"'dt = Me v' ()...) d)'" = Mv (I) e. (9.11.10)


o 0

Here v (I) denotes the value of the relative velocity of the gas in the outlet
cross-section (it equals zero in the inlet).
Equation (4) now takes the form

M [~O +w x vo + w x r~ + w x (w x r~)] = V + <P + F Car . (9.11.11)

Here <I> denotes the thrust

<P = -eMv (I). (9.11.12)

Its value is IMlv (I) = -Mv (I) and its direction coincides with that of e
as M < O. The effect of the motion of the gases in the nozzle results also
in the Carialis force
F Car = -2w x eMI. (9.11.13)
What remains is to make the corresponding transformations in eq. (5).
The position vector r' of a point on the nozzle axis can be represented in
the form

r' = r~ + e (I - )...) , (9.11.14)


where r~ denotes the radius vector of the point on the axis of the outlet
cross-section with the origin at the pole 0 of axes Oxyz fixed in the rocket.
As shown above
*'
r dm = eMd)"', r
. **' .
dm = Mev' ()...)d).... (9.11.15)
9.12 Gyroscopic platform 515

Hence

J
I

= 2Mr; x (w x e) l + 2Me x (w x e) (l-)..) d)"


o
= 2M (r; + ~el) x (w x e) l
= - (r; + ~el) x F CoT (9.11.16)

as well as

J J J
I I

r'x **'
r dm M r; x ev' ()..) d)" + M (l-)..) ex ev' ()..) d)"
o 0
Mv(l)r; x e = -r; x <1>. (9.11.17)

We arrive at the equation of rotation of the rocket

eO . w+ w x eO . w + Mr~ x (vo +w x vo)

rno + r; x <1> + (r; + ~el) x F CoT . (9.11.18)

It has the form of the equation of motion of a free rigid body. The right
hand side is completed by the moments of the thrust <1> applied to the
outlet cross-section and Coriolis's force applied at the mid-point of the
nozzle about pole O.
Relationships (11) and (18) present the required equations of motion of
the rocket. The assumptions that the gas motion is stationary and the gas is
incompressible are essential for the derivation. Additional terms appearing
when these effects are accounting for are usually small and can be neglected,
see [28].
The equation of motion could also be derived by using the equations of
dynamics of variable mass by Meshersky, see [56].

9.12 Gyroscopic platform


The carrying body is a platform with mounted gyroscopes which are gim-
bals with rotors. The axes Oxyz with the unit vectors h, i 3, i3 are fixed in
the platform, the angular velocity vectors and the velocity of pole 0 are
denoted by wand Vo, respectively. An orthogonal trihedron of unit vectors
ak, bk, Ck = ak x bk is bound to each gimbal, vector ak being directed along
the rotation axis of the k - th gimbal and has a constant direction with
respect to axes Oxyz bound to the platform. The direction of vector b k
516 9. Dynamics of relative motion

coincides with that of the corresponding rotor. The angles of rotation of


the gimbal relative to the platform are denoted by Xk and 'Pk, then the
angular velocity vectors of these bodies with respect to axes Oxyz are

(9.12.1)

see Fig. 4.3 taking into account the change in notation. The derivatives of
vectors ak, bk, Ck referring to this coordinate system are given by

(9.12.2)

Hence

(9.12.3)

The point Ok at the intersection of the rotation axes of the gimbal and the
rotor is assumed to coincide with the centres of inertia of these bodies, r~
denoting the position vector OO~. The masses of the k - th gimbal and
rotor are denoted by m~ and m%, respectively, and their inertia tensors at
point Ok are as follows

B/Ok = A~E + (C~ - AU bkb k , B"Ok = A%E + (C~ - A%) bkb k ,


(9.12.4)

where, for brevity,

Ak = A~ + A%, Ck = C~ + C~, mk = m~ + m%. (9.12.5)

We proceed to construct the equations of motion for the platform which


is the carrying body. Equations (6.6) and (6.15) are applicable, the number
of carried body being 28 (8 gimbals and 8 rotors).
-=--=-> .
Let m and r~p = OCp denote respectIvely the platform mass and the
radius vector of its centre of inertia. Then eq. (6.6) describing the motion
of the centre of inertia for the system of bodies under consideration has the
form

m [~o +w x Vo + W x r~p + wx (w x r~p)] +


s
L mk [~o +w x Vo + w x r~ + wx (w x rU] = V.
k=1

The terms rk *' are omitted in eq. (6.6) since the centre of inertia Ok does
not move with respect to axes Oxyz. The pole 0 is taken in the centre of
inertia of the platform and masses mk at points Ok, then
s

mr~p + Lmkr~ = 0 (9.12.6)


k=l
9.12 Gyroscopic platform 517

and the equations of motion for the centre of inertia is

M (~·o +w xVO) = V, (9.12.7)

where M is the mass of the system. By the appropriate attachment of bal-


ance masses one can make point 0 coincident with the point of intersection
of the axes of Cardan's suspension, the latter being used for mounting the
platform of the moving base, e.g. ship, airplane etc.
Let us turn to the equation of rotation (6.15). The tensor of inertia of
the platform and masses mk at point 0 is given by
s
eO = e~ + I: mk (Er~ . r~ - r~r~) . (9.12.8)
k=l

Using equalities (3), (6) and notation (5), (8) we arrive at the equality

s
w Xbkw· b k) = rn O - I: {AkakXk + [Ckw x ak+
k=l

2 (Ck - Ak) Ckbk . w] Xk + C~ (0kbk + CkCPkXk + w x bkCPk)}. (9.12.9)

Let us proceed to the equations for the carried bodies. In principle we can
use eq. (6.39) but Lagrange's equations is a more effective way to construct
the equations.
The kinetic energy of the k - th gimbal with the rotor is equal to

n= ~ (w + akXk) . [EA~ + (C~ - A~) bkbk] . (w + akXk) +


1
"2 (w + akXk + bkCPk) . [EA% + (C~ - AD bkbk] . (w + akXk + bkCPk)·
Transformation yields

(9.12.10)

Provided that there exist no active forces, the generalised force QXk is
the moment of the frictional force of the k - th gimbal, whilst Q'Pk denotes
the difference between the torque and the moment of the frictional force of
the k - th rotor. In what follows we assume

Q Xk = 0, Q'Pk = 0, (k = 1, ... ,s). (9.12.11)


518 9. Dynamics of relative motion

Coordinates 'Pk are cyclic. However Xk are the positional coordinates, as


vectors b k are dependent on them. We can construct s equations which
express that the cyclic momenta P'!'k (the kinetic moment of the k - th
rotor) are constant. The result is
(9.12.12)
By means of eq. (7.17.2) and (7.15.4) we construct the Routhian function

Rk ~ [Ak (X~ + 2Xkw, ak + w 2) + (C~ - AU (w· bk)2 +

2Hkw· b k - ~~] (k=l, ... ,s) (9.12.13)

and the differential equations


Ak (Xk + w· ak) - (C~ - A k) W· bkCk . W - Hw· Ck = 0
(k = 1, ... , s). (9.12.14)
While deriving the latter equation the following relationship

ob k *
0 bk
- - = - - =Ck
OX oX
was used. Now it is easy to prove eq. (9) which is the vectorial form of the
Euler-Lagrange equation for the Routhian function
1 s
R= 2w.eO ·w+ LRk. (9.12.15)
k=l

The first of these equations


d oR oR oR
- - - +W2- -W3-- =ml
°
dt OWl OW3 OW2
takes the form
s
(eO. wf + W2 (eO. W)3 - W3 (eO. W)2 + L [Ak (Xkakl + WI) +
k=l

(C~ - A k) (w· bkbklt + Hkhkl + AkXk (W2ak3 - W3ak2) +


(C~ - A k) w . b k (W2bk3 - W3bk2) + Hk (W2bk3 - W3bk2)] = mf.
It represents the projection of the vectorial relationship

(eO + tEAk) . W + w x eO . w + t {AkakXk + AkXk w x ak+

(C~ - A k) [w· bkbk + W· bkW X b k + (w· Ckbk + W· bkCk) Xk+


Hk (XkCk +w x bd]} = mO. (9.12.16)
9.12 Gyroscopic platform 519

on axis Ox. Equation (9) can be used to deduce this form if <Pk and <Pk are
replaced by the corresponding expressions from eq. (12) and the following
equality

is applied.
The rotors are assumed to rotate with high angular velocities <Pk whereas
the angular velocity of the platform and the angular velocity Xk of the pro-
cession remain sufficiently low, so that only terms with the kinetic moments
of the rotors are kept in eq. (14). Then we have

W· Ck =0 (k = 1, ... , s). (9.12.17)

If the number of gyroscopes s 2: 3 and there are three non-coplanar vectors

Ci . (Ck x Cl) =I 0 (i, k, l = 1, ... , s), (9.12.18)

the consequence of equalities (17) is that the angular velocity of the plat-
form becomes zero

W=o, (9.12.19)

that is the platform is stabilised in space. Equation (16) yields


s s
LHkCkXk = rno or LHkbk x akXk +rno = 0 (9.12.20)
k=l k=l

and expresses the" equilibrium equation" of the platform, that is the ge-
ometric sum of the gyroscopic moments of the rotors and the resultant
moment of the external forces acting on the platform about its centre of
inertia equals zero.
As an example let us consider the case of three gyros mounted on the
platform in the way shown in Fig. 9.6.
Figure 9.7 displays the orientation of the unit vectors iI, i 2, i3, vectors
ak of the axes of the gimbal rotation and the initial positions of vectors
b k , Ck. It is easy to see that

al = h, b l = -il cos Xl - i2 sin Xl' cl=ilsinXl-bcoSXl' }


a2 = -b, b 2 = il cos X2 + i3 sinX2, C2 = -~l sin X2 + ~3 c~s X2,
a3 = i3 , b 3 = -il sin X3 +b cos X3, C3 = -11 cos X3 - 12 sm X3·
(9.12.21)

Condition (18) takes the form

cos (Xl - X3) cos X2 =I O. (9.12.22)


520 9. Dynamics of relative motion

.,
I
I
I

~MI

o.g.
I Mz
L__________ J
FIGURE 9.6.

It is not met at Xl = X3 ± 7r /2, i.e. when the rotation axes of the first and
third rotors become parallel to each other as well as at X2 = 7r /2, i.e. when
the rotation axis of the second rotor becomes perpendicular to the axes the
first and third rotors.
It presents no problem now to write down the equations of motion of
the platform and the gyroscopes in expanded form. For instance, the first
equation in (14) is

A1 (Xl + W3) - (C~ - A 1) [W1W2 cos 2X1 + ~ (W~ - wi) sin 2X1] -

H1 (W1 sin Xl - W2 cos Xl) = O. (9.12.23)

Taking the rotation angles as well as the projections of the angular velocity
of the platform on the coordinates axes as small quantities yields

A1 (Xl +W3) +H1W2 = 0, }


A2 (X2 - W2) - H2W3 = 0, (9.12.24)
A3 (X3 + W3) + H3W1 = O.
Within this approximation we reduce the three equations obtained by
projecting the vectorial relationship (16) on axes h, h, h to the form
9.12 Gyroscopic platform 521

~
00
1:,
1 b,o
"l
:;lei CD

III
~,

FIGURE 9.7.

QlWl - H3 (W3 + X3) = mf , }


Q2 W2 - HI (WI + Xl) = m~, (9.12.25)
Q3 W3 - H2 (W2 - X2) = mf,
where

Ql = ef + C~ + C~ + A3, Q2 = e~ + C~ + AI , Q3 = ef + A 2·
(9.12.26)

Now we can cast the systems of equations (24) and (25) as follows

.. H§ 1 .a
WI + A3Ql WI = Ql ml ,
.. Hi _ 1 .a (9.12.27)
W2 + Al Q2 W2 - Q2 m2 ,
. Hi 1 .a
W3 + A2Q3 W3 = Q3 m3 .

Under the assumption that the projections of the resultant moment rna
on the platform axes are constant, the projections of the angular velocity
vary harmonically with the frequencies
H3 HI H2
)'l = y'A3(h ' A2 = ~' A3 = JA2Q3 ' (9.12.28)

which are very high since they are proportional to the angular velocities of
the rotors. We have

WI = Ml sin (Alt + ad, W2 = M2 sin (A2t + 0"2), W3 = M3 sin (A3 t + (3)


(9.12.29)

and, by virtue of eq. (25) ,

. mg .( )
Xl = - HI +M3 sm A3t + a 3 - V~M2COS A2t+a2
M ( ) (9.12.30)

etc. The angles of rotation of the gimbals increase, however small the mo-
ments m~ may be. In order to avoid this effect special facilities are used.
522 9. Dynamics of relative motion

One of them, pursuing also other aims, is schematically depicted in Fig.


9.6. Three motors M I , M 2 , M3 produce torques depending on the angles of
rotation of the gimbals, [43].
10
Canonical equations and Jacobi's
theorem

10.1 Legendre's transformation


Let us consider a continuous function <I> (Xl, ... , xn) of variables Xl, ... , Xn
which has continuous derivatives of the first and second order with respect
to all variables. The transformation from the "old" variables Xl, ... , Xn to
the "new" ones

0<1>
Ys =- (8 = 1, ... ,n) . (10.1.1)
oXs

is carried out by means of function <I> referred to as the generating function.


The Jacobian of the variables y is the Hessian H (<I» of function <1>. Its
elements are the second partial derivatives of <I> with respect to Xs

0 2<1> 0 2<1> 02<1>


Ox2I OX20XI OXnOXI
H(<I» = (10.1.2)
0 2<1> 0 2<1> 02<1>
OXIOX n OX20Xn OX~

The system of equations (1) can be resolved for the old variables if the
Hessian is non-singular. Then we can obtain the formulae for the inverse
transformation by expressing the old variables in terms of the new ones

Xs = XS (Yl, ... ,Yn) (8 = 1, ... ,n) . (10.1.3)


524 10. Canonical equations and Jacobi's theorem

This transformation can be written in the form of eq. (1). To this end, the
generating functions of the new variables \[1 is introduced as follows
n
\[1 (YI,'" ,Yn) = L XkYk - «I (Xl, ... ,Xn ), (10.1.4)
k=l
where all old variables are replaced in accordance with eq. (3). Indeed,
taking the partial derivative of \[1 with respect to Ys and noticing that \[1
depends on this variable, both explicitly and in terms of the old variables,
we obtain

since the sums cancel out by virtue of equalities (1). Thus


0\[1
Xs = - (s = 1, ... ,n), (10.1.5)
oYs
as required. The relationships (1) and (5) defining the direct and the inverse
transformations are referred to as Legendre's transformation. It is easy
to understand that applying inverse transformation twice results in the
recovering of the original transformation.
If, in particular, the generating function «I is a quadratic form

(10.1.6)

then its Hessian is the determinant lal of the matrix a of the coefficients of
this form

H(<(I)=lal· (10.1.7)

Provided that a is a non-singular matrix, i.e. its determinant lal is not zero,
then the system of linear equations

0«1 n
Ys = ox = Laskxk (s = 1, ... ,n) (10.1.8)
s k=l

can be solved for the old variables. Denoting the inverse matrix by b = a-I
we find
,n
Xs = LbskYk (s = 1, ... ,n). (10.1.9)
k=l
It follows from Euler's theorem on homogeneous functions that the gener-
ating function \[1 (YI,'" ,Yn) of the inverse transformation is obtained by
10.1 Legendre's transformation 525

direct substitution of expressions (9) into eq. (6). Indeed, by means of eqs.
(1) and (4) we have

(10.1.10)

where <I> should be expressed in terms of the new variables. Thus

(10.1.11)

i.e. the quadratic form \I! obtained by means of matrix b a-I is the
associate expression for the quadratic form <I>. It is easy to understand that
\I! can be written in the form of the following determinant

au aI2 ain YI
1 a21 a22 a2n Y2
\I! = - - (10.1.12)
21al an2
anI ann Yn
YI Y2 Yn 0
An example of the application of Legendre's transformation is the gener-
alised Castigliano theorem. Let us consider an equilibrium configuration of
a system with ideal constraints subject to active forces of two types: poten-
tial forces prescribed by the potential energy II (ql, ... , qn) and the applied
loads F I, ... , F N. The sum of the elementary work of these forces due to
virtual displacements of the system points from the equilibrium position
must vanish

Q: denoting the generalised forces of the loads. Provided that the variations
8q8 are independent of each other we obtain the equilibrium equations

all
Q8* = Oq8 ( 1 , ... ,n)
S = (10.1.13)

which can be regarded as the transformation defining the new variables,


which are the generalised forces of the loads, in terms of the old variables
which are the generalised coordinates q8 of the system in the equilibrium
position under the above loads. The potential energy II is the generat-
ing function of this transformation. The generating function of the inverse
transformation R (QI, ... ,Qn) is the quantity
N
R(Ql, ... ,Qn) = LQ8q8 -II, (10.1.14)
8=1
526 10. Canonical equations and Jacobi's theorem

referred to as the complementary work. Relationships (5) are as follows


8R
qs = 8Qs (8 = 1, ... ,n). (10.1.15)

They have more general meaning as those given by Castigliano's theorem in


Sec. 5.7. The latter expresses the above result for the special case in which
the potential energy is a quadratic form of the generalised coordinates. Ac-
cording to eq. (11) the complementary work R is the associate expression
for the potential energy and Castigliano's theorem acquires the conven-
tional formulation. The generalised Castigliano theorem has a practical
significance in those cases when the complementary work can be explic-
itly expressed in terms of the generalised forces. This remark is also true
in regard of any Legendre transformation which in no way simplifies the
solution of the system of equations (1) but simply indicates the possibility
of their representation by means of the generating function expressed in
terms of new variables.

10.2 Canonical equations of motion


Let us consider a material system with n degrees of freedom subject to
holonomic constraints. Lagrange's equations of this system present a sys-
tem of n second order differential equations for the generalised coordinates
resolved for the second derivatives iis as pointed out in Sec. 7.4. There exist
a number of ways to replace this system of equations by a system of 2n
first order equations. For instance, it is sufficient to denote qs = r sand
to introduce into consideration 2n variables qs, rs which are independent
variables determining the behaviour of the system. The system of 2n first
order equations of motion consists of n Lagrange's equations expressing rs
in terms of qs, rs and n equations qs = rs. Of course, this trivial change
of notation is immaterial for the solution. The above-said pursues simply
the objective to indicate that the consideration of 2n first order differen-
tial equations instead of n second order equations allows us to view all 2n
variables as being independent of each other. An example of the fruitful
application of first order equations is the concept of the Euler-Lagrange
equations since the latter allows us, under an appropriate choice of the
quasi-velocities, to obtain a more transparent and symmetrical form of the
equations of motion than Lagrange's equations.
We can adopt the generalised coordinates ql, . .. , qn and the generalised
momenta PI, ... ,Pn (which are linear functions of the generalised velocities
qI, ... ,qn) as the 2n variables of the system of differential equations of
motion. The generalised momenta are defined in Sec. 4.2 as follows
8T
Ps = 8qs (8 = 1, ... ,n). (10.2.1)
10.2 Canonical equations of motion 527

Under such a choice of variables and for the potential active forces, the
equations of motion are written down in a very compact and symmetrical
form referred to as the canonical form. This simplifies the analysis of the
general properties of the motion and reduces the problem of integration of
the canonical equations to a search for the complete integral of the equa-
tion with first order partial derivatives (Jacobi's theorem). The variations
qs,Ps are independent and are symmetric in the forthcoming equations and
transformations.
The kinetic energy, regarded as a function of the generalised velocities,
plays the part of the generating function transforming the old variables iJs
to the new ones Ps. The Hessian of the transformation is the determinant
of matrix A composed of the coefficients of the quadratic form T2 which
is the part of T. As shown in Sec. 4.2 this determinant does not vanish
as it is positive and for this reason equations (1), linear in the generalised
velocities, iJi are solvable. This results in the relationships (4.2.6)
(10.2.2)
which are linear in the generalised momenta. They comprise the first set
of the canonical equations, namely a system of n first order differential
equations defining the time-derivatives of the generalised coordinates in
terms of the generalised momenta, generalised coordinates and time.
Legendre's transformation enables eq. (2) to be represented in another
form. In order to obtain this form, let us construct, by means of eq. (1.4),
the generating function of the inverse transformation
n
T = LPsiJs - T = T (qt, ... , qn,P1,··· ,Pn; t), (10.2.3)
s=l
in which the generalised velocities are replaced by their expressions (2) in
terms of the generalised momenta. Then, by virtue of eq. (1.5), we obtain

.
qs = -aT (s = 1, ... ,n) . (10.2.4)
oPs
It is important that the second set of equations of motion is constructed by
means of the same function T. In contrast to equations (4), which simply
express the transformation, the equations of the second set express the
general equation of dynamics. The latter is preferable to take in the form
of Lagrange's fundamental equations (6.4.6)

(10.2.5)

or in the form (3)


d n n _ n
dt L Psl5qs = l5 L PsiJs - l5T + L Qsl5qs. (10.2.6)
s=l s=l s=l
528 10. Canonical equations and Jacobi's theorem

We obtain

or

The last group of terms cancel out due to relationships (4), whereas the next
to last cancels out since variation and differentiation are interchangeable
(the rule db = 8d). As mentioned above, this rule could be avoided if
we would apply the general fundamental equation (6.4.11) instead of the
Lagrange fundamental equation. We arrived at the same result

Because the variables 8qs are independent we find

.
Ps=-~+Qs
aT (s=l, ... ,n). (10.2.7)
uqs

This is the second set of the canonical equations for the system of variables
qs,ps. It will coincide with the second set of Hamiltonian equations provided
that the forces are potential. Similar to Lagrange's equations, which can
be obtained by means of a single function, namely the kinetic potential
L, a single function H referred to as the Hamiltonian function suffices to
construct the canonical system. This function is defined as follows
n
H (qI, ... ,qn,P1,··· ,Pn; t) = T + II = LPsqs - T + II (10.2.8)
s=l

or
n
H = L Psqs - L = T + II. (10.2.9)
s=l

Function H depends on the generalised coordinates, generalised momenta


and time, whilst the generalised velocities are removed by means of rela-
tionship (2).
It is worth noticing that time t may appear explicitly in the Hamiltonian
function in terms both of T (under non-stationary constraints) and the
generalised potential energy II.
10.2 Canonical equations of motion 529

Taking into account that the potential energy does not depend on the
generalised momenta we obtain, instead of eqs. (4) and (7),

. at aH . at all aH
qs=-=- Ps=----=--·
aps aps' aqs aqs aqs
We derive the system of canonical Hamiltonian equations

. aH
Ps=-- (8 = 1, ... ,n). (10.2.10)
aqs
Variables qs and Ps satisfying the system of canonical Hamiltonian equa-
tions are called canonical.
If the constraints are stationary then T is a quadratic form of the gen-
eralised velocities and t
is the associate expression for the kinetic energy
t
according to the previous section. As is denoted by T' in Sec. 4.2 we have
under stationary constraints

H = T' + II, (10.2.11)

that is the Hamiltonian function is the total mechanical energy expressed


in terms of the generalised coordinates and generalised momenta. By virtue
of eq. (1.12) construction of T' reduces to calculation of the following de-
terminant

A11 A12 Aln PI


A21 A22 A 2n P2
, 1
T = -21AI (10.2.12)
AnI An2 Ann Pn
PI P2 Pn 0

As mentioned in Sec. 4.1, under non-stationary constraints,


n
LPsqs = 2T2 + Tl,
s=1
and it follows from eq. (9) that
n
H = LPsqs - L = 2T2 + Tl - (T2 - Tl + To - II) = T2 + (II - To) ,
s=1
(10.2.13)

where T2 should be expressed in terms of the momenta.


In the case of stationary constraints, T' and thus H are quadratic forms
of momenta. When the constraints are non-stationary, T2 is a quadratic
form of the generalised velocities. However the latter are linear forms in the
generalised momenta with free terms. Hence, H consists of the terms which
530 10. Canonical equations and Jacobi's theorem

are quadratic and linear in the momenta and a free term. The corresponding
t
expressions for are given by formulae (4.2.13) and (4.2.14). In eq. (4.2.13)
i' designates the difference T2 - To expressed in terms of the generalised
velocities and coordinates whereas, in eq. (4.2.14), T' stands for the same
difference in terms of the momenta.
Let us construct the total time-derivative of the Hamiltonian function
according to the canonical equations (10)

dH 8H n (8H. 8H.)
dt {it +~ 8qs qs + 8ps Ps
8H
8t +
t
8=1
(8H 8H 8H 8H) 8H
8qs 8ps - 8ps 8qs - 8t' (10.2.14)

Hence the Hamiltonian function H retains a constant value throughout the


motion provided that H does not contain time explicitly. We obtain the
integral of the equations of motion

(10.2.15)

which is termed the energy integral. For a system with stationary con-
straints it states that the sum of the kinetic and potential energies is con-
stant.
If the coordinates qm+h ... , qn are cyclic, then the equalities

. 8H
Pm+l = - - - - (l = 1, ... , n - m) , (10.2.16)
8qm+1
which are consequences of the definition of Sec. 7.15 and the Hamiltonian
function (13), can be used to conclude that the corresponding generalised
momenta remain constant throughout the motion

Pm+l=i3m+l (l=I, ... ,n-m). (10.2.17)

In this case, the system of 2m first order equations


. 8H
Ps=-- (8 = 1, ... , m) (10.2.18)
8qs
with the same number of unknown variables, which are the positional coor-
dinates and the corresponding momenta, is under consideration. Assuming
that this problem is solved then the determination of the cyclic coordinates
from the equations
. 8H 8H
qm+l = --- = --- (l = 1, ... , n - m) (10.2.19)
8Pm+l 8(3m+l
is reduced to quadratures since the right hand sides of these equations
become given functions of time.
10.3 Explicit form of the canonical equations 531

10.3 Explicit form of the canonical equations


In Sec. 7.8 we related the motion of a particle subject to stationary con-
straints to the motion of the representative point on the Riemannian man-
ifold with the metric defined by the square of the linear element
1
ds 2 = Tdt2 = 2aa{3dqadq{3,

and introduced the velocity vector v (7.8.4) of this point with the covariant
components equal to the generalised velocities ga. The relation between
covariant and contravariant components is given by the general rule (B.2.3)
and in this case takes the form
.{3 _ aT _
aa{3q - aga - Pa· (10.3.1)

Thus, under the scaling of Sec. 7.8, the generalised momenta represent the
covariant components of the velocity of the representative point. From eq.
(1) we obtain the inverse relationships

(10.3.2)
representing the first set of the canonical equations. The associate expres-
sion for the kinetic energy is given by

(10.3.3)

Taking into account that the expression for the potential energy does not
contain the momenta, we arrive at another form of equations (2)
.a aT' a (T' + II)
q =- (10.3.4)
aPa aPa
The second set of the canonical equations represents the law of motion
(7.7.11), that is "the acceleration of the representative point equals the
acting force", in the covariant form
all
Wa = Qa = - aqa·
Here, by analogy with eq. (7.8.8), we differentiate the velocity vector v =
Para using rule (B.4.19) to obtain

Wa = Pa - { : , } P{3g'Y = Pa - a'Y6 { : , } P{3P6,


and

(10.3.5)
532 10. Canonical equations and Jacobi's theorem

The latter implies an expanded expression of the second set of the canonical
equations. In order to prove this let us consider the equality
&T' 1 &a{3"f
&qo. = "2 &qo. P{3P6·

By recalling Ricci's theorem (B.5.8) and the rule of covariant differentiation


(B.5.6) we can write

V' a{3"f = &a{3"f


a &qo.
+{ (3 } aP"f
ap
+{ 'Y
ap
} a{3p =0
,

so that

-~ ({ :p } aP"f +{ :p } a{3p) P{3P"f

(3 } &T'
- { ap aP'Y P{3P"f = - &qo. '

and comparing with eq. (5) yields

. & (rr + T') &H


Po. = - &qo. - &qo. '

as required. Thus, the explicit form of the canonical equations is as follows

q.o. = ao.{3p{3, p. _ a"f6 {


a -
(3 } P P
a8 {3 "f -
orr
&qo.· (10.3.6)

10.4 Examples
10.4.1 Motion of a particle in a central force field
The kinetic and potential energies, in terms of the spherical coordinates
R, {j, A, are given by expressions (7.18.6). We obtain

(10.4.1)

The Hamiltonian function will be

H = T' + rr = 1 (Ph
-2
m
+ RP~ + R /.12
sm {j
) + rr (R) . (10.4.2)

As pointed out above, the first set of the canonical equations is nothing
else than equations (1) resolved for the generalised velocities

R= &H =PR . &H P>.


A= - = 2 . (10.4.3)
apR m' &p>. mR2 sin {j
lOA Examples 533

The second set of equations is

(10.4.4)

Thus, the system has the energy integral and the cyclic integral P>.. = const.

10.4.2 Canonical equations of motion of a heavy top under


given motion of the support point
Figure 10.1 shows the system of "half-moving" unit vectors n, n' , i3 intro-
duced in Sec. 2.10. The velocity vector of the centre of inertia of the top is
equal to

Vc = Vo + (~n + ~sin -On') x i~zc = Vo + Zc ( -~n' + ~sin -On) ,


with Vo denoting the velocity vector of the support point O. We obtain

2 2+ +zcP(.0v 2+ 'ljJ.2sm. 2)
Vc = Vo , (.0
-0 + 2zc -vvo· n ' +
· ·'ljJ .sm0vVo . n ) ,

and the kinetic energy is given by

1 (.2
T = '2A -0 + 'ljJ·2 sin2)
-0 + Mzc (--ovo·
. n' + 'ljJsin-ovo·
. n) +

'12 C ( rp + 'ljJcos-o
. )2 + '2Mv5,
1 (10.4.5)

where A and C designate the equatorial and the axial moments of inertia
at point O.
The generalised momenta are

P1j; A~ sin2 -0 + Mzcvo . nsin -0 + C (rp + ~ cos-o) cos-o,


PJ9 A~-Mzcvo·n', p<.p=C(rp+~cos-o).
From these equations we obtain

1;; = .12 (P1j; - P<.p cos-o - Mzcvo· nsin -0),


Asm -0
.0
v= 1 (PJ9+Mzc, v o·n') ,
A (10.4.6)

. P<.p -
cp=-C cos
·2
.0 M'
-0 (P1j;-P<.pCOSv- . .0)
zcvo·nsmv.
Asm -0
534 10. Canonical equations and Jacobi's theorem

FIGURE 10.1.

By means of eq. (2.113) we obtain the Hamiltonian function

H = ~ 2 (p,p-Pcpcos{)-Mzcvo·nsin{))2+ (10.4.7)
2Asm {)
2
1 ( , ,)2 1 Pcp ( , {) ) 1 2
2A Pt1+ Mz c v o· n +2C+ Mg zccos +(0 -2 Mvo ·
Here (0 (t) denotes the height of the support point. The terms

Mg(o - 21 Mv o2
in the expression for H are immaterial.
While constructing the second set of the canonical equations, we should
consider that vectors nand n' are functions of angles 'l/J and {)

n i1cos 'l/J +hsin 'l/J, n' = (-hsin 'l/J +i 2 cos'l/J)cos{)+i3 sin{),


i~ - (-h sin'l/J + i2 cos'l/J) sin {) + is cos{),

where h , i 2, is are unit vectors fixed in the space. Then we obtain the
following equalities

on on'
o'l/J
8i 3 , ::l"
vl3 . {)
(10.4.8)
o{) = 0, = -ncos{), o{) = -n, o'l/J =nsm .

Taking them into account, we have

P,p - oH
o'l/J = M Zc [.'l/Jvo· o n+·'l9 (vo . n) cos {)]
o'l/J sin-a
Mzc (vo· nsin{)t - Mzcwo· nsin{),
10.4 Examples 535

where Wo = Vo stands for acceleration of the pole O. The second equation


reduces to the form
oH cos7'J , 2
>l~Q • 3 (p,p - P<p cos 7'J - M zcvo . n sin 7'J) -
uu Asm 7'J
A P.<P
Slnu
~a (N - P<p cos 7'J - M z~vo . n sin 7'J) +
, . ,(. . on')
Mgzc sm 7'J + M Zc vo· mj; cos7'J - 7'J v o' o7'J .

The last terms can also be cast as follows

-Mzc
' (vo· on' . on' .)
o'ljJ 'ljJ + vo· o7'J 7'J =
,(
-Mzc vo· n
')_ + Mzcwo'
, ,
n.

Entering the quantities


P,p* = P,p - M zcvo'
' . ~a
nsmu, p~ = P19 + Mz~vo' n',
we can recast the canonical equations as follows
1) the first set

;p = .1 2
Asm 7'J
(p~. -
'f'
P<p cos 7'J) ,
. 1
7'J = AP~, (10.4.9)

. P<p
'P = C - cos 7'J (*
A sin 2 7'J P,p - P<p cos
7'J)
;
2) the second set
P,p
.* = - M zcwo'
' . u,
n SIll ~Q )

p~ = A Si~3 7'J (p~ - P<p cos 7'J) (p~ cos 19 - p<p) + (10.4.10)
Mgz~sin7'J + Mz~wo' n',
p; = o.
We notice that due to eq. (8)
• ~Q 0 .,
Wo . nSlnu = o'ljJ Wo . 13,

Therefore, in full accordance with Sec. 9.2 and formula (9.2.9), the motion
of the support point can be considered by including the additional term
no = M' .,
zCW O' 13= M wo·rc
'

into the expression for the potential energy. Equations (9) and (10) are the
canonical equations for the Hamiltonian function

H* 1 ( * 19) 2 1 *2
2A sin 2 7'J P,p - P<p cos + 2A P19 +
1P<p
2C 2+M ' cos u~a
gzc + M'
zcwo . 1./3, (10.4.11)
536 10. Canonical equations and Jacobi's theorem

where

Wo . i~ = (io sin 'ljJ - Yo cos'ljJ) sin f) +.20 cos f), (10.4.12)

xo, Yo, Zo denoting the projections of the acceleration of pole 0 on axes


fixed in the space.

10.4-3 Canonical equations of motion of a heavy top carrying


a flywheel
It is assumed that the bearings of the flywheel's axis lie on the axis of
symmetry Oz' of the top. Denoting the rotation angle of the flywheel by
X, we set

a = f3 = 0, 1=1

r
in eq. (4.12.3) which yields the formula for the kinetic energy

T = ~ A (l + l sin 2 f)) + ~C (cp + ~ cos f) +


~83X2 + 8 3X (cp + ~ cos f)) . (10.4.13)

Here A and C denote the equatorial and the axial moments of inertia of the
top and mentally "frozen" flywheel about axes passing through the support
point of the top, and 8 3 is the moment of inertia of the flywheel about its
axis of rotation. Then we obtain

pfJ A~, P'1f; = A~ sin2f) + C (cp + ~ cos f)) cos f) + 8 3 Xcos f),
P<p C (cp + ~ cos f)) + 8 3 X, Px = 8 3 X+ 8 3 (cp + ~ cos f)) ,
and

(10.4.14)

These equations comprise the first set of the canonical equations. The
Hamiltonian function equals

H = "21 [p~
A +
(N - P<pcosf))2
A sin2 f)
1 (2 C 2
+ C - 8 3 P<p + 8 3Px - 2P<PPx
)]
+Qz~cosf), (10.4.15)

where Q and z~ denote the weight and the abscissa of the centre of inertia
of the system, respectively.
10.5 The Poisson brackets and the Lagrange brackets 537

We can add four integrals to the first set of the canonical equations,
namely, the integral of energy

H=h (10.4.16)
and the three cyclic integrals

(10.4.17)
The problem is reduced to quadratures, the distinction from the problem
of the heavy top being only in the determination of angle cpo Angle X is
obtained immediately

(10.4.18)

10.5 The Poisson brackets and the Lagrange


brackets
We consider two functions u and v of two sets of independent variables,
each set containing n variables

U=U(ql, ... ,qn,PI,··· ,Pn) =u(p!q), }


(10.5.1)
V=V(ql, ... ,qn,PI,··· ,Pn)=v(p!q).
Let us construct the expression

8u 8v 8u 8v (8U 8v 8u 8V)
8qI 8PI + ... + 8qn 8pn - 8PI 8qI + ... + 8pn 8qn '

referred to as the Poisson bracket of these functions and denoted by

(u v) =
,
t(
k=1
8u ~ - ~~)
8qk 8Pk 8Pk 8qk .
(10.5.2)

This definition yields immediately the following properties of a Poisson


bracket

(u,v)=-(v,u), (u,u) =0, (u,c) =0, (10.5.3)


provided that c does not depend on the variables under consideration. Also

(10.5.4)
The partial derivatives of function u with respect to qs and Ps, can be cast
in terms of the Poisson brackets as

(10.5.5)
538 10. Canonical equations and Jacobi's theorem

which allows us in particular to write the Hamiltonian canonical equations


in the form

(10.5.6)

Finally, let us notice some values of the Poisson brackets

(10.5.7)

with 8 st denoting the Kronecker delta. These brackets are known as the
fundamental brackets.
If it is necessary one can use subscripts to indicate the independent
variables, i.e.

(u,v)q,p.

In what follows we use Jacobi's identity, namely

(u(v,W)) + (v, (w,u)) + (w, (u,v)) = O. (10.5.8)

In order to avoid cumbersome notation we notice that each term on the


left hand side of this identity contains the second derivative of one of the
functions u, v, w. The second derivatives of w appear in the first two terms
but not in the third term. According to eq. (3) these first two terms can
be written as follows

(u (v, w)) - (v, (u, w)). (10.5.9)

The brackets (v, w) and (u, w) can be cast as linear forms in the first
derivatives of w as

(v,w) = =:t
B(w)
k=l
(Bk;w
Pk
+ Bn+k;w)
qk
, )

:t
(u,w) = A (w) =
k=l
(Ak ;w
Pk
+ An+k ;w)
qk
,
(10.5.10)

where

au
An+k = - apk . (10.5.11)
10.5 The Poisson brackets and the Lagrange brackets 539

Then

(u (v, w)) - (v, (u, w)) = A [B (w)]- B [A (w)]

Ln(
s=1
Bsa + B n+s 8 L Aka + An+ka
a
Ps
a)n(
qs k=1
ow
Pk
ow)
qk

= L L (AsBk - AkBs) aa
n n [a2w
s=1 k=1 Ps qk
+ (AsBn+k - An+kBs) a
a 2w
a +
Ps qk

(An+sBk - Bn+sAk) a
a2aw + (An+sBn+k - a2 w ]
An+kBn+s) ~ + ...
qs Pk uqsuqk
where dots denote the terms containing no second derivatives of w. The
expression under the double sum reverses the sign under a change of sum-
mation indices s to k and k to s. Hence this sum is equal to zero and the
second derivatives of w in eq. (9) cancel out. Repeating this reasoning for
functions u and v, we find that the second derivatives of u and v must van-
ish. However, each term contains a derivative of one of the three functions,
thus all terms cancel out in pairs which completes the proof.
Another property of the Poisson brackets is used below, namely, if u and
v depend also on time t, then

(10.5.12)

This can be proved easily by means of the definition of the Poisson bracket.
Let us proceed to the expression

referred to as the Lagrange bracket and denoted by

[ ] = ~ (aqk aPk _ aqk aPk) (10.5.13)


u, v ~ au av av au .
k=1
Particularly, by analogy with eq. (7), we have

These brackets are called the fundamental Lagrange brackets. When needed
we will use the independent variables q, P as subscripts.
Let

(10.5.15)
540 10. Canonical equations and Jacobi's theorem

denote n pairs of the independent variables


(10.5.16)
We can take quantities (16) as functions of the variables (15). Let us con-
sider the 4n2 Poisson brackets
(Uk,U s ), (Uk,V s ), (Vk,U s ), (Vk,V s ), (k,s= 1, ... ,n), (10.5.17)
which are elements of the quadratic Poisson matrix
(UI, UI) (UI,U n) (UI,VI) (UI' vn)
........ . ........ . ........ . .........
(un, UI) (Un, Un) (un,vd (Un,vn)
P=
(VI, UI) (VI, Un) (VI, VI) (VI, Vn)
........ . ......... ........ . . ........
(V n, UI) (V n, Un) (Un,UI) (Un,u n)
(10.5.18)
see (A.1.28).
By analogy, we can introduce the matrix A whose elements are the La-
grange brackets
(10.5.19)
These skew-symmetrical matrices can be written in the form of 2 x 2 block
matrices

P _II ((u,u))
- ((v,u))
((u,v))
((V,V))
I , A-II
-
[[u,u]]
[[v,u]]
[[U, v]]
[[V, v]]
I, (10.5.20)

where each block denoted by ((u, u)), ((u, v)), ... , [[v, v]] is a n x n matrix,
see (A.1.23) and (A.1.28). Notice that matrices P and A reverse sign under
transposition
pI = _ P, A = -A, (10.5.21)
which follows from the properties of the Poisson bracket and the Lagrange
bracket.
The product AP is considered in Appendix A, where the following iden-
tity is proved
PA = AP = -E2n, (10.5.22)
see eq. (A.2.39). The latter equation is equivalent to the four matrix equal-
ities
((u,u)) [[u,u]] + ((u, v)) [[v,u]] = -En, }
((u,u)) [[u,v]] + ((u, v)) [[v,v]] = 0,
(10.5.23)
((v,u)) [[u,u]] + ((v,v)) [[v,ull = 0,
((v,u)) [[u, v]] + ((v,v)) [[v,vll = -En'
10.6 Poisson's theorem 541

For instance, an expanded form of the first equality is


n
L {(Us,Uk) [Uk,Ur ] + (Us,Vk) [Vk,Ur ]} = -8 sr . (10.5.24)
k=l

10.6 Poisson's theorem


Let us consider function rp of 2n variables q, p and time t. The complete
derivative constructed due to the Hamiltonian canonical equations (2.10)
equals

which, using the Poisson brackets, can be recast in the form

<P = ~~ + (rp, H). (10.6.1)

Thus, if

rp = rp (qlPi t) (10.6.2)

is an integral of the equations of motion, i.e. rp is constant throughout the


motion, then
arp
at + (rp, H) = o. (10.6.3)

This equality is satisfied for all arguments q,p, t, i.e. it is an identity.


Based upon Jacobi's identity and relationship (5.12) it presents no prob-
lem to prove Poisson's theorem: if

(10.6.4)

are two integrals of the canonical equations, then the Poisson bracket

(10.6.5)

has constant value c throughout the motion.


Indeed, according to eq. (3) we have the identities

(10.6.6)

and it is necessary to prove that


542 10. Canonical equations and Jacobi's theorem

By virtue of eqs. (5.12) and (6) this relationship is reduced to Jacobi's


identity

- (('PI' H) , 'P2) - ('PI' ('P2' H)) + (('PI' 'P2) , H)


('P2' ('PI' H)) + ('PI' (H, 'P2)) + (H, ('P2' 'PI)) = 0,
which completes the proof.
The quantity 'IjJ is either a new integral of the canonical equations, or a
function of known integrals, in particular, integrals (4)
(10.6.7)
or a constant which can also be zero.
Let us consider the case of H which does not contains time t explicitly.
Then, there exists the energy integral (2.15). Combining it with the integral
(2) we obtain, by virtue of Poisson's theorem
(H, 'P) = a,
where a is constant. Making use of equality (3) we find the new integral
a'P
at = -a. (10.6.8)

Continuing our reasoning we arrive at the next integral


a 2 'P
at 2 = al

etc. Provided that 'P = c is an integral containing no time t explicitly, then


expression (H, 'P) which, due to eq. (3), is identically equal to zero, yields
no new integral.
We notice that combining integrals (3.4) corresponding to the cyclic co-
ordinates with one of integrals (3.5)

'P(qI, ... ,qm;PI,··· ,Pm;(3m+ I ,··· ,(3n;t) =c, (10.6.9)


we arrive at the identity

(10.6.10)

yielding no new integral.


Jacobi [44] referred to Poisson's theorem as "one of the most remarkable
theorems of the whole of integral calculus. In the particular case when
H = T + II it is the fundamental theorem of analytical mechanics". In
order to obtain a new integral by combining some integral with a given
one, as Jacobi suggested, it is necessary for this new integral to be an
integral belonging to the problem under consideration. However the first
integrals are obtained from the general principles, for instance the law of
conservation of area, thus they do not belong to the problem in question
and it is not feasible to require that they yield all of the first integrals.
10.7 Canonical transformations 543

10.7 Canonical transformations


Let us consider two systems of variables, the old variables

(10.7.1)

and the new variables

(10.7.2)

which are considered as functions of the old variables and time t

Qs ==- Qs (ql, ... ,qn;.PI, ... ,Pn;.t) : Qs (qIP: t), }


Fs - Fs (ql, ... ,qn,PI,··· ,Pn, t) - Fs (qlp, t) (10.7.3)
(s=l, ... ,n).
These relationships are assumed to be solvable for the old variables, i.e.

(10.7.4)

The transformation (3) is called canonical if the following condition is


met: the following expression
n n
LPk t5qk - L Fi t5Qi (10.7.5)
k=1 i=1

is a variation of a certain function of the old and new variables and time
t. The time is not varied, so that varying the first set of relationships (4)
we have

(10.7.6)

and expression (5) can be represented in the form

(10.7.7)

The transformation will be canonical if the three conditions are satisfied:


firstly, in eq. (7) the derivative of the coefficient of t5Qi with respect to Ql
must be equal to the derivative of the coefficient of t5Ql with respect to Qi
544 10. Canonical equations and Jacobi's theorem

secondly, the derivatives of the coefficients of 8Pi and 8Pl with respect to
~ and Pi must be equal to each other

and thirdly the derivatives of the coefficients of 8Qi and 8~ with respect
to PI and Qi must be equal to each other

Now taking into account that the new variables Pi, Qi are independent, i.e.

8Pi =0
8Ql '

and recalling the definition ofthe Lagrange bracket (5.13), we arrive at the
equalities

[Ql, Qi] = 0, [PI, Pi] = 0, [Qi, Pzl = 8il (i, l = 1, ... ,n), (10.7.8)

under which transformation (3) will be canonical. These conditions are


necessary and sufficient. The property that the transformation is canonical
is reversible, that is if transformation (3) is canonical, then the inverse
transformation (4) is canonical too. This follows from condition (5) in which
the old and the new variables are on equal terms.
Equalities (5.14) for the fundamental Lagrange brackets, in terms of new
variables Pi, Qi, can be cast in the form

(10.7.9)

Thus, instead of eq. (8) we can write the relationships

(10.7.10)

expressing the invariance of the fundamental Lagrange brackets under the


canonical transformation.
In accordance with eq. (8), the Lagrange matrix A of the system of
functions Q s, P s has the form

(10.7.11)
10.7 Canonical transformations 545

where 0 denotes the n x n null matrix. By means of eq. (5.22) it is easy to


conclude that the Poisson matrix has the same structure, i.e. the following
equalities

hold along with eq. (8). They present other forms of the necessary and
sufficient conditions of the canonical transformation (3). The fundamental
Poisson bracket (5.7) is also the invariant of the canonical transformation.
The following statement is of a more general character: functions u, v
are considered first as depending upon the old variables, then as depending
upon the new variables related to the old ones by means of the canonical
transformation. Then

(u, v)qp = (u, v)QP. (10.7.13)

This can be proved by direct differentiation. Indeed, by expressing in the


Poisson bracket

the partial derivatives with respect to the old variables by the partial deriva-
tives with respect to the new ones by the formulae

inserting into the bracket we obtain, after summation over k,

(u, v)qp

In the case of the canonical transformation, that is under condition (12)


we can write

(u, v)qp ~ ~ (8U 8v 8u 8V)


~~ 8Qr 8Ps - 8Pr 8Qs Drs

~(8U 8v 8u 8V)
~ 8Qr 8Pr - 8Pr 8Qr = (u,v)QP'
546 10. Canonical equations and Jacobi's theorem

which completes the proof. The invariance of the Lagrange bracket under
the canonical transformation, i.e.

[u, vl qp = [u, vl Qp (10.7.15)

is proved by analogy.

10.8 Generating functions


Let us return to the definition of the canonical transformation. The function
whose variation is given by expression (7.5) referred to as the generating
function can depend on all 4n variables qs,Ps,Qs,Ps and t. However in
as much as there exist relationships (7.3) and (7.4) it suffices to consider
it to be dependent on 2n arguments and time t, among them n old and n
new arguments. Thus we have four types of generating function

(10.8.1)

In accordance with the choice of generating function, it is preferable to


write the condition (7.5) for the canonical transformation in one of the
forms
n
L (PiOqi - PiOQi) = oV (q, Q; t), I (10.8.2)
i=1

- t (qiOPi + PiOQi) = 0 [VI (q, Q; t) - t qiPi] = oV3 (p, Q; t),


(10.8.4)

n
-L (qiOPi - QiOPi) o [VI (q,Q;t) - tqiPi + tQiPi]
i=1
oV4 (p, P; t) . (10.8.5)

From these equations we obtain the following systems of canonical trans-


formations
aV
I
Pi=~, (i=l, ... ,n), (10.8.6)
uqi
10.8 Generating functions 547

oV2 Q•. -_ oV2 ( )


(10.8.7)
Pi = oqZ' OPi i = 1, ... ,n ,

p. __ oV3 ( )
(10.8.8)
•- OQi i = 1, ... ,n

Q•. -_ oV4
OPi
(
i = 1, ... ,n .
)
(10.8.9)

Let us consider, for instance, transformation (6). Taking a function

and constructing formulae (6), we should resolve the first set of equations
(6) for Qi which is needed for obtaining expressions (7.3) relating the new
variables in terms of the old ones. The solution exists when the Jacobian
does not equal zero, that is

(10.8.10)

This is the requirement imposed on the generating function Vi, Expressions


for the new variables Q s in terms of the old ones qi, Pi obtained from the
first set of equations (6) should be substituted into the second set of these
equations. This yields the second set of equalities (7.3) expressing Ps in
terms of the old variables.
The conditions for the other generating functions can be written down
by analogy. For instance

02V2 02V2
OP1Oql oP1oqn
#0, (10.8.11)
02V2 02V2
OPnOql oPnoqn
and the procedure for constructing formulae for transformation (7.3) re-
duces to determining the new variables P s from the first set of equations
(7) and inserting these expressions into the second set of these equations.
The equalities relating the generating functions are given by eqs. (2)-(5).
It is function V2 which is considered in what follows more frequently than
548 10. Canonical equations and Jacobi's theorem

the others. Its expression in terms of the other generating functions has the
form

n
VI (q, Q; t) = V2 (q, P; t) - L QiPi, (10.8.12)
i=1

n
V3 (p, Q; t) = V2 (q, P; t) - L (QiPi + qiPi) , (10.8.13)
i=1

n
V4 (p,P;t) = V2 (q,P;t) - LqiPi. (10.8.14)
i=1

Notice that if V2 (q, P; t) is considered as the generating function of Legen-


dre's transformation from q, P to p, Q, then - V3 is the generating function
for the inverse transformation from p, Q to q, P. For the partial transfor-
mations of only one set of variables to other, i.e. from q to P and from Q to
P, carried out by the generating function 112, the inverse transformations
are due to - V4 and -VI, respectively.
Relationships (6)-(9) yield the following equalities

8Pi 8Qs
(10.8.15)
8Ps 8qi '

(10.8.16)

Using these equalities let us find the Jacobian of the canonical transforma-
tion [16]

By virtue of the property of Jacobians we can write

D (QIP) = D (QIP) : D (qlp) . (10.8.17)


qlp qlQ D (qIQ)
10.8 Generating functions 549

On the other hand,

OQ1 OQ1
Oq1 oqn 1 o
oQn oQn 0 1
Oq1 oqn
D (QIP)
qlQ 1 1
Oq1 oqn 0 o
oPn oPn 0 0
Oq1 oqn
(-It D (PI' ... ,Pn ) .
q1,··· ,qn
By analogy we also obtain

and the first relationship in eq. (15) takes the form

D (QIP) = D
qlQ
(~)
qlQ
. (10.8.18)

It follows from eq. (17) that the Jacobian for the transformation (Q, P) --+
(q,p) is equal to unity. It is known however that the products of the direct
and the inverse transformation is also equal to unity, thus

D(QIP)
qlp
=D(~)
QIP
=l. (10.8.19)

The consequence of this statement is Liouville '8 theorem on the invariance


of the integral

J... J
2n
dq1 ... dqndp1 ... dPn

over a certain volume v of the space of the canonical transformation (q, p),
i.e. the phase space, under the canonical transformation (q, p) --+ (Q, P).
Under any transformation

J... J
2n
dq1 ... dqndp1 ... dPn = J... J
2n
D (~:~ ) dQ1 ... dQn dP1 ... dPn .

Here the integration on the right hand side is carried out over volume V re-
sulting from v under this transformation. For the canonical transformation
550 10. Canonical equations and Jacobi's theorem

(19), we have

J... J
2n
dqI ... dqndPI ... dPn = J... J
2n
dQI ... dQndPI ... dPn·

(10.8.20)

10.9 Invariance of the canonical transformations


Let us recall that the variables satisfying the system of the canonical equa-
tions, i.e. the generalised coordinates qs and generalised momenta Ps, are
referred to as the canonical variables. The fundamental importance of the
canonical transformations in analytical mechanics is due to the theorem
that variables Qs, P s related to the canonical variables qs,ps by a canon-
ical transformation are also canonical, that is, they satisfy the system of
canonical Hamilton's equations
. 8K . 8K
Qs = 8Ps ' Ps = - 8Qs· (10.9.1)

Here K denotes the Hamiltonian function depending on the new canonical


variables Qs, Ps and time t which is related to the Hamiltonian function of
the old canonical variables qs,Ps as follows

K =H 8l1; (10.9.2)
+ at'
lI; being the generating function of the canonical transformation in ques-
tion. In particular, if the latter does not depend explicitly on t, then
K=H. (10.9.3)
Let us consider the transformation with generating function ~ (q, Q; t).
The variables qs, Qs are considered as independent, and applying this trans-
formation we have
8Qs =0 (10.9.4)
8t .
On the other hand, we obtain from the second set of formulae (8.6)
8Ps 8 2 VI
8t - 8t8Q s .
(10.9.5)

Let us consider now Qs, Ps as functions of time and the old canonical
variables prescribed byeq. (8.6). Then, based upon relationships (6.1), (4)
and (5), we can write down the equalities

(10.9.6)
10.9 Invariance of the canonical transformations 551

which, due to the invariance of the Poisson brackets under the canonical
transformation (7.13) and the property (5.5), can be cast in the form

(10.9.7)

What remains is to notice that

since V1 does not depend on Ps . Denoting

K=H BV1
+ Bt '
we arrive at the system of the canonical equations (1)

. BK
Qs = BPs' (s=l, ... ,n),

which is required. Nothing would be changed in this derivation if the gen-


erating function V3 was used.
In the case of the generating function V2 (or V4 ) we should take

BPs = 0
Bt

instead of eqs. (4) and (5). This results in

since Qs does not appear in expression V2 (q, P; t). Taking

K = BV2
Bt +H,
we arrive at the canonical equations.
Thus it is proved that the canonical equations are retained under the
canonical transformation of variables.
552 10. Canonical equations and Jacobi's theorem

10.10 Examples of canonical transformations


10.10.1 First example
For the function of the old coordinates and new momenta
n
V2 (q, P) = L Psis (q1, ... ,qn), (10.10.1)
s=1

which is independent of time t, condition (8.11) takes the form

(10.10.2)

When it is satisfied, V2 can be taken as the generating function of the


canonical transformation, and the second set of equalities (8.7)

(10.10.3)

describes the point transformation of the generalised coordinates. Expres-


sions for the new momenta Ps in terms of the old generalised coordinates
and momenta, due to the first set of the above equalities, are obtained from
the system of linear equations

n als
Pi = LPs~ (i = 1, ... ,n), (10.10.4)
s=1 q.

which has a solution since its determinant (2) is non-zero.


For example, let us consider motion of the particle in the field of central
force using the cylindrical coordinates z, r, >.. The kinetic and potential
energies are given by

and the Hamiltonian function is

(10.10.5)

In order to transform to spherical coordinates we should consider the func-


tion

(10.10.6)
lD.lD Examples of canonical transformations 553

which can be viewed as the generating function of the canonical transfor-


mation since condition (2) is met. We obtain by means of eq. (3)

Q1 = J z2 + r2 = R, Q2 = arctan ~ = {),
z
Q3 = A, (10.10.7)

R, {), A denoting the spherical coordinates. By virtue of formulae (4) we


find

and the Hamiltonian function, in terms of the new variables, is obtained


by substitution of these expressions into eq. (5)

H = 2~ (P{ + ~2 pi + ;2 Pi) + f (R)

2~ (p~ + ~~ + R2 :i~2 {)) + f (R) , (10.10.8)

where PR,P1'J,P).. are the new momenta corresponding to the spherical co-
ordinates. This expression for H is given in example 10.4.1.

10.10.2 Second example


Let us consider the transformation
8F
Qs = qs, Ps = Ps - ~, (10.10.9)
uqs
in which F is a function of ql, ... , qn and t. To prove whether this trans-
formation is canonical, let us construct expression (7.7) for the Poisson
brackets. We have

By analogy

and
~ (8Qs
(Qs,Pr ) =~ 8Pr 8Qs 8Pr ) ~
aa-a~ =~8sk8rk=8sr.
k=1 qk Pk Pk uqk k=1
This transformation is seen to be canonical. The generating function can
be taken in the form
n
V2 =L qsPs + F (ql, . .. ,qn; t) . (10.10.10)
s=1
554 10. Canonical equations and Jacobi's theorem

10.11 Canonical equations of the relative motion


In accordance with eqs. (4.9.1), (4.9.5) and (4.9.6), the kinetic energy of
the system is as follows

1 21
T = -Mvo+-w·e 0 ~or~.
·w+M ( voxw).rc+Mvo·~-;;-qs+
'
2 2 ~1u~

(10.11.1)

where r~ (qt, ... ,qn) denotes the position vector of the particles with re-
spect to the pole 0 of the coordinate basis fixed in the carrying body. Its
motion, described by the velocity Vo of the pole 0 and the angular velocity
vector w, is assumed to be given. The position vector of the centre of inertia
referring to the pole 0 is denoted as r~. Finally, Tr is the kinetic energy of
the relative motion which is a quadratic form of the generalised velocities.
The expressions for the momenta are

oT oTr ~,or~ or~


Ps = ~ = ~ +w· ~miri x a + Mvo · 8 ' (10.11.2)
qs qs i=1 qs qs
and the Hamiltonian function, due to eq. (2.13) is equal to

H = Tr + IT - To = Tr, + IT - '2w·
1 e 0 . w + M ( w x Vo ). rc,
' (10.11.3)

1
since the term '2Mv5 in (1) depends only on time and can be omitted for
this reason. Replacing in the expression for Tr the generalised velocities in
terms of the momenta we obtain T;. Let us enter matrix B = A -1, which
is the inverse matrix for the matrix A of the coefficients of the quadratic
form T r . Then we have, due to eq. (2),

T'r

(10.11.4)

Here the quantities


or~ 0 ,
Ps = Ps - Mvo . - - = Ps - -Mvo . rc (8 = 1, ... ,n) (10.11.5)
oqs oqs
are introduced. Adopting that

Qs = qs (8 = 1, ... ,n) , (10.11.6)


10.11 Canonical equations of the relative motion 555

then the transformation relating the new variables Ps , Qs to the old ones
Ps, qs will be canonical of the type (10.9) with

F = Mvo· r~ = M (VOlX~ + V02Y~ + V03Z~). (10.11. 7)

The partial derivative of F with respect to time is equal to

at
8F
= M (. ,
VOlXe
. ,
+ V02Ye . Z ') M
+ V03 e =
* ,
Vo ·re, (10.11.8)

and the Hamiltonian function K for the system of the canonical variables,
due to eqs. (9.2) and (3) is

K = T: + IT + ITw + Mr~ . (~·o +w x v o) = T: + IT + ITw + nO.


(10.11.9)

Here, according to notation (9.2.10) and (9.2.9), nw and no denote the


potential energy of the centrifugal forces and the inertial forces of the pure
translation, respectively. The value of T:
is given by eq. (4).
The system of the canonical equations of the relative motion is written
in the form

(s=I, ... ,n), (10.11.10)

the previous notation for the generalised coordinates being adopted here.
The action of the centrifugal forces and inertia forces caused by the
translational motion is taken into account by terms ITW and ITo in the
expression for the potential energy. The Coriolis forces and the rotational
forces of inertia, being non-potential, are automatically accounted for by
introducing the new momenta Ps . However it would be an error to identify
them with the momenta under the relative motion

(r) _ 8Tr
Ps - 8. (8 = 1, ... , n). (10.11.11)
qs

The associate expression for the kinetic energy in terms of these quantities
can be represented by the quadratic form

T' =
r
~2 ~
~~
~ B s kP(r)p(r)
s k'
(10.11.12)
s=lk=l

and, by comparison with eq. (4), yields

N 8'
Ps =p(r)
s +w· ~m·r'
~ t t
x -..!.i
8 (8 = 1, ... , n) , (10.11.13)
i=l qs
556 10. Canonical equations and Jacobi's theorem

so that

Here the differentiation rule

was used. Noticing that

w· w x[ (, ar~)] = 0,
ri x aqs

and, recalling (9.2.14), we can recast eq. (14) in the form

. _
Ps - Ps
'(7") _ W
Qs +W L m,
• N
i=l
~
. (""
'"
uqk
~
x '"""
uqs
,
+ r, x '" "'2') .
_u_r_i_
'"
uqkuqs
(10.11.15)

The right hand side of the second set of the canonical equations (10) has
the form
aT;
--- - -
an
- -- ---
anw anD
(10.11.16)
aqs aqs aqs aqs .
While taking the derivative of T; given by eq. (4) it is necessary to take
into account that both Bsk and the following sum
N '" ,
"""' ' uri
~miri x 8'
i=l qs
depend on qs, so that by virtue of eqs. (4), (13) and the first set of equations
in (10), we have

(10.11.17)
10.12 Canonical transformation and the process of motion 557

The first term represents the derivative of T:


given by eq. (12) and the
derivative is taken under the assumption that the relative momenta are
independent of the generalised coordinates.
Recalling the notation of Sec. 9.2, the second set of equations (10) be-
comes

. (r)
Ps = QW
s + 2w . ~
L mar~
i- ar~. - -
x --qk - - - a (IT + ITW
a*T: + ITO)
i=l aqs aqk aqs aqs
or

aqs (IT + IT
aqs _ ~
s = _ a*T: + ITO) + QWs + QCor
P·(r) w (
s 8 = 1, ... , n ) ,
(10.11.18)

whilst the first set, due to eq. (12), is given by the equalities

. = ---c;:)
qs aT: (8 = 1, ... , n ) . (10.11.19)
aps
The system of equations (18) and (19) can be constructed based upon
the equations of motion (2.4) and (2.7) and considering only the relative
motion with account of the inertia forces being determined by the motion
of the coordinate basis Oxyz.
The system of equations (10), obtained by means of the canonical trans-
formation of the canonical equations of the absolute motion, is an inter-
mediate form between the latter and the equations of the relative motion.
Because of the terms ITW and ITo in the Hamiltonian function this system
takes into account the inertia forces of the potential character whereas the
other inertia forces are not explicitly set off and this allows one to retain
the Hamiltonian form of these equations.

10.12 Canonical transformation and the process of


motion
It is assumed that the general solution of the system of canonical equations
. aH (8 = 1, ... ,n) (10.12.1)
Ps = - aqs

is known. This solution yields expressions for the generalised coordinates


and momenta as functions of time and 2n arbitrary constants denoted by
ak and i3 k

qs =qs(t,al, ... ,an;i31'··· ,i3n), (10.12.2)


Ps =ps(t,al, ... ,an;i31,··· ,i3n), (8=1, ... ,n). }
558 10. Canonical equations and Jacobi's theorem

The system of these equations must be solvable for constants ak and (3k,
in other words, we require a non-zero Jacobian

D (ql, ... , qn;Pl,··· ,Pn) # 0, (10.12.3)


(al, ... ,an ;(3l,··· ,(3n)
otherwise solution (2) is not general since not all variables are independent
of each other.
We write down the solution of the system of equations (2) for constants
ak and (3k in the form

as =as(t,ql, ... ,qn;Pl,··· ,Pn),


(10.12.4)
(3s =(3s(t,ql, ... ,qn;Pl,··· ,Pn)' (s=l, ... ,n). }
These expressions provide us with all 2n integrals of the canonical system
of equations (1). Hence, according to Poisson's theorem any of the Poisson
brackets

(10.12.5)

is either constant (in particular equal to zero identically) or is expressed


in terms of the same constants ai, (3i; otherwise the brackets (5) would
contain the (2n + 1) - th independent integral which is impossible.
Assume nOw that we took not an arbitrary system of 2n independent
integrals of the canonical systems of equations (1) but its total Cauchy's
integral. This implies that constants ak and (3k are defined as the values
of the generalised coordinates and momenta at time instant t = to

(qs)t=to = as, (Ps)t=to = (3s (s = 1, ... ,n).


Then, recalling the definition (5.13) of the Lagrange's bracket, we obtain,
for example,

and thus

By analogy we find

[as, ak]t=to = 0, [(3s, (3d t =to = o.


Based on these equations and relationships (5.22) (or (5.24)) we obtain
similar expressions using Poisson's brackets
10.12 Canonical transformation and the process of motion 559

However, as mentioned above, all Poisson's brackets retain constant values,


hence the relationships obtained must hold for any t. Therefore,

(10.12.6)
which means that relationships (2) representing the Cauchy integral of
the canonical equations are formulae for the canonical transformation of
variables OOk,f3 k to qs,Ps. The inverse transformation is also canonical.
By virtue of the above said the motion of the holonomic system is a
canonical transformation of the values of the generalised coordinates and
momenta at time instant t = to to the values at the current time instant t.
In this sense, the motion is said to be a progressively developing canonical
transformation.
It appears that the question as to what meaning has the generating
function of the canonical transformation has, is of great importance. We
return to this point in Sec. 12.8.
In closing we consider two illustrative examples.
1) The system of equations

q = a cos A (t - to) + ~ sin A (t - to) , } (10.12.7)


p = -OOA sin A (t - to) + f3 cos A (t - to)
is the Cauchy integral of the system of the canonical equations with the
following Hamiltonian function

(10.12.8)

Equations (7) describes the motion of an oscillator. Let us prove, by calcu-


lating Lagrange's bracket, that equations (7) present a canonical transfor-
mation. We have
oq op oq op 2 . 2
[a, f3l = 000 0 f3 - 0 f3 000 = cos A (t - to) + sm (t - to) = 1,

which is required since this is the only Lagrange's bracket in the case of a
single-degree-of-freedom system.
2) In the problem of parabolic motion of a heavy particle of unit mass in
a vacuum the expressions for the coordinates x, y and momenta x, y have
the form

x = Xo + Xo (t -
to), Y = Yo + Yo (t - to) - -21 9 (t - to )2,} (10.12. 9)
x=xo, y=y-g(t-to)·
These equalities are Cauchy's integrals of the system of canonical equations
with the following Hamiltonian function
1
H = "2 (x 2+ y2) + gy. (10.12.10)
560 10. Canonical equations and Jacobi's theorem

Let us prove that transformation (9) is canonical. As follows from the


table of the type (5.18), it is necessary to find altogether ~ (4n2 - 2n) = 6
Lagrange's brackets

[xo, Yo], [xo,Xo] , [xo,zio],


[Yo, xo], [Yo, Yo] ,
[xo, yo].

A simple calculation shows that all conditions (7.8) are fulfilled. For in-
stance,

and so on.

10.13 Jacobi's theorem


It is proved in Sec. 10.9 that under the canonical transformation

Qs = Qs (qlp; t) , P s = P s (qlp; t), (s = 1, ... , n) (10.13.1)

the new variables Ps , Qs satisfy the system of Hamiltonian equations

. 8K . 8K
Qs = 8Ps ' Ps = - 8Q s ' (s = 1, . .. , n) , (10.13.2)

obtained by means of the Hamiltonian function

8V
K=H+7jt. (10.13.3)

Here H denotes the Hamiltonian function for the initial canonical variables
8H
Pk=-- (k=l, ... ,n) (10.13.4)
8qk

and V is the generating function. Clearly, K should be expressed in terms


of the new canonical variables P s , Qs'
Thus, given a generating function V, we reduce the problem of integra-
tion of the initial system of equations of motion (4) to the new problem
(2) which may be simpler than the first. There arises a natural question
regarding the choice of generating function V for which K = 0 or what is
immaterial K = const. Then, by virtue of eq. (2), P s and Qs are constants

Qs=a s , P s =(3s (s=l, ... ,n). (10.13.5)


10.13 Jacobi's theorem 561

For this reason, if we seek a generating function of the type V2 (q, (3; t)
depending on the old coordinates qs and new momenta (3 s' then its deter-
mination, due to eq. (3), reduces to the solution of the Jacobi-Hamilton
equation

(10.13.6)

which is a partial differential equation. In this equation, in accordance with


the transformation formulae (8.7) Pi are replaced by the partial derivatives
of the sought-for function V with respect to the generalised coordinates

Pk =-
oV (k = 1, ... ,n). (10.13.7)
Oqk

The complete integral of the Jacobi-Hamilton equation (6) has n constants


(31' ... , (3n· Additionally, it must satisfy condition (8.11) which has the form

02V 02V
0(31 0 ql o(31 oqn
(10.13.8)
02V 02V
o(3n o ql o(3n oqn

under the adopted notation. In particular, if at least one of the n constants


appears in V additively, i.e. if

then the Jacobian (8) is zero and the solution is not the required complete
integral.
Having the complete integral of the Jacobi-Hamilton equation, we arrive
at the system of equalities by means of (8.7)

(k=l, ... ,n). (10.13.9)

The first of these systems is solvable for ql, ... ,qn by means of eq. (8).
The solution yields expressions for ql, ... ,qn in terms of t and 2n con-
stants as, (3s. Inserting these expressions into the second system, we find
the momenta as functions of the same values t, as, (3s

(10.13.10)

We obtained the complete integral of the canonical system of equations (4)


since it contains 2n independent constants.
562 10. Canonical equations and Jacobi's theorem

Thus we proved Jacobi's theorem: the problem of obtaining the com-


plete integral of the canonical system of equations (4) is equivalent to the
problem of constructing the complete integral of the Hamilton-Jacobi equa-
tion (6) in partial derivatives. The complete integral is the solution of the
Hamilton-Jacobi equation containing n arbitrary constants and satisfying
the condition (8) of non-zero Jacobian. The general integral of the canoni-
cal system and the system of finite equations (9) are obtained by means of
this complete integral.
Let us notice that the general integral of the canonical system (10),
obtained by Jacobi's theorem, is a canonical transformation of variables
as, (3 s to variables qs, Ps carried out with the help of the generating function
V. Any other general integral of the form

(10.13.11)

Ck , Dk being arbitrary constants, does not possess this property and does
not allow one to obtain the complete integral of the Jacobi-Hamilton equa-
tion since relationships (11) do not present a canonical transformation.
According to the theorem of Sec. 10.12 the exception is the case when Ck
and Dk are equal to the initial values of the coordinates and momenta.
As proved in the theory of differential equations, the problem of con-
structing the complete integral of the Hamilton-Jacobi equation and the
general integral of the canonical system are mathematically equivalent.
Generally speaking, they are equally difficult. However we can mention a
number of special cases when the Hamilton-Jacobi equation can be solved
more easily than the canonical system. This point is the subject of Sec.
10.14. More importantly, the solution (10) obtained by means of Jacobi's
theorem is a canonical transformation and this fact, as it is shown in Chap-
ter 11, simplifies essentially the form of equations of the perturbed motion.
The case of the existence of the energy integral when time t does not
appear explicitly in the Hamiltonian function is of special interest. Then
we can fulfill the Hamilton-Jacobi integral (6) by setting

(10.13.12)

Substitution of this expression into eq. (6) leads to the Hamilton-Jacobi


equation for function W

H (q!, ... ,qn; ~: ' ... , ~:) = h. (10.13.13)

It is obtained by replacing the momenta in the expression for the energy


integral H = h by the partial derivatives of the sought-for function W with
respect to the corresponding coordinates.
10.13 Jacobi's theorem 563

The problem is now reduced to the search for the complete integral of
eq. (13) containing, apart from h, n - 1 constants

W=W(ql,'" ,qn;/31,'" ,/3n-l;h) (10.13.14)

and satisfying the condition

#0. (10.13.15)
o{3n-l o ql o/3n-l o qn
02W 02W
oh8ql ohoqn
We can construct now equations (9) which take the form
oW
o{3k = Cl:k (k = 1, ... , n - 1) , (10.13.16)

oW
oh = t - to, (10.13.17)

oW
Pk=- (k=I, ... ,n). (10.13.18)
Oqk
The constants h and to play the part of {3n and Cl: n .
Equations (16) are n - 1 dependences between the generalised coordi-
nates ql, ... , qn and do not contain time. They produce a set of all trajec-
tories in the space of these variables of dimension n depending on 2n - 1
variables. The feasibility of explicitly obtaining the trajectories, avoiding
the conventional procedure of removal of t from the equations of motion,
is a remarkable feature of Jacobi's method.
It is easy to see that at least one minor determinant corresponding to the
element of the last row of Jacobian (15) is not zero, otherwise inequality
(15) does not hold true. We can take

Equations (16) yield then the parametric representation of the trajectories

qs=qs(qn;Cl:l, ... ,Cl:n-l;{31,'" ,(3n-l,h) (s=I, ... ,n-l),


(10.13.19)
564 10. Canonical equations and Jacobi's theorem

qn playing the role of the parameter. Inserting these expressions into the left
hand side of eq. (17) results in function of qn only. Resolving this equation
for qn and making use of eq. (19) we arrive at the equations of motion

qs=qs(t-to;a1, ... ,an-1;,61, ... ,,6n-1,h) (8=1, ... ,n).


(10.13.20)

Of course they can be obtained by solving the system of equations (16)


and (17), whose solvability follows from condition (15). Substitution of
expression (20) into the right hand sides of expression (18) yields

Ps = Ps (t - to; a1,··· , a n-1; ,61,··· , ,6n-1, h) (8 = 1, ... , n).


(10.13.21)

Relationships (20) and (21) represent the canonical transformation of the


variable ,6k, h, fYk, -to to the variables qs, Ps determined by the generating
function

v = -ht + W (q1, ... , qn, ,61,··· , ,6n-1, h) (10.13.22)

of the type V2 of coordinates q1, ... , qn and new momenta ,61,··· , ,6n-1' h.
Provided that qm+1, ... , qn are cyclic coordinates, the Hamilton-Jacobi
equation (13) has the form

(10.13.23)

and its solution should be sought in the form

(10.13.24)

Substitution into eq. (23) leads to the Hamilton-Jacobi equation for func-
tion W*

8W* 8W* )
H ( q1,···,qm, 8q1 '···'8qm;,6m, ... ,,6n-1 =h. (10.13.25)

The complete integral of this equation contains, apart from n - m con-


stants ,6m, ... , ,6n-1, h, other m -1 constants ,61, ... , ,6m-1· The following
condition
8 2 W* 8 2 W*
8,61 8q1 8,61 8qm

8 2 W* 8 2 W* #0 (10.13.26)
8,6m-1 8q1 8,6m-1 8 qm
8 2 W* 8 2 W*
8h8q1 8h8qm
10.13 Jacobi's theorem 565

must hold. The function of the type V2 is taken as the generating function.
Let us proceed to the generating function of the type VI. In order to main-
tain consistent notation ak and (3 k for the coordinates and momenta, we
cast the complete integral of the Hamilton-Jacobi (6) as follows

(10.13.27)

Then, by means of eq. (8.6), we obtain, instead of (9), the following system
of equations

(k=l, ... ,n), (10.13.28)

where, by virtue of eq. (8.10), condition (8) in which (3k is replaced by ak


must hold.
For the generating functions of the type V3 we will obtain, due to eq. (5),

V3 (pIQ; t) = V3 (PI, ... ,Pn; 001,··· ,an; t). (10.13.29)

Then, by virtue of eq. (8.8)

(s=l, ... ,n), (10.13.30)

and the Hamilton-Jacobi equation will take the form

(10.13.31)

Here the generalised coordinates among the arguments of the Hamiltonian


function H are replaced by the partial derivatives of the required gener-
ating function with respect to the momenta. If we deal with an arbitrary
canonical system prescribed by an arbitrary function H there is no reason
to prefer equation (6) to eq. (31). This is not the case for the canonical sys-
tem of equations of dynamics since the Hamiltonian function is a quadratic
function of momenta Ps and an arbitrary function of the coordinates. For
this reason, the generating function V3 , as well as V4, is of little use, however
exceptions are possible.
Returning to eq. (14) let us ascribe the constant values of (31, ... ,(3n-l, h,
then the relationship

(10.13.32)
I being a parameter, can be deemed as an equation for the family of hy-
persurfaces of the Riemannian manifold Rn. Notice that superscripts are
used in the latter equation since tensor analysis notation is applied. On the
surface I = const

oW {) s = 0 (10.13.33)
oqS q ,
566 10. Canonical equations and Jacobi's theorem

where the variations 8q s are the contravariant components of the infinites-


imal displacement vector 8r on this surface, so that 8r = r s8 q s, r S are the
basis vectors and 8q s = r S • 8r. Recalling equalities (18) we can write eq.
(33) in the form
oW
~r
S s:
. ur S
= Psr .
s:
ur = o. (10.13.34)
uqs

Thus, the momentum vector (or the velocity vector of the representative
point)
(10.13.35)
is orthogonal to the surface of constant Wand, by virtue of eq. (18), is
equal to the gradient of this function
p = gradW (10.13.36)
This equality becomes especially illustrative in the case of motion of a free
particle in the potential field. Then
p = mv = grad W. (10.13.37)
The trajectories of the particle are orthogonal to the surfaces W = const.

10.14 Separability of variables in the


Jacobi-Hamilton equation
In some cases the complete integral of the Hamilton-Jacobi equation can
be obtained for the Hamiltonian functions of special form which admit the
separation of variables. In these cases the complete integral is represented
as a sum of functions, each depending only on one of the generalised coor-
dinates.
We consider the case of stationary constraints. The associate expression
for the kinetic energy is then a quadratic form of the generalised momenta.
It is assumed that this form contains only the momenta squared and their
products are absent. Then the Hamilton-Jacobi equation (13.13) takes the
form

2"L
1 n (OW)2 +II(ql, ... ,qn)=h.
A k(ql, ... ,qn) a (10.14.1)
k=l qk
The question arises as to with what structure of functions Ak and II does
this equation have the complete integral of the form
(10.14.2)
having the necessary number of constants.
To begin with, let us consider a few examples.
10.14 Separability of variables in the Jacobi-Hamilton equation 567

10.14.1 Keplerian motion


Assuming that the trajectory of the particle in the field of the attracting
force to a fixed centre is a plain curve, we will describe the position of the
particle in this plane by two polar coordinates rand 'P. The expression for
the Hamiltonian function for a particle of unit mass is

(10.14.3)

According to eq. (13.24), the complete integral of the Hamilton-Jacobi


equation

should be sought for in the form

(10.14.4)

where R (r) is defined by the following differential equation

[R' (r)]2 = 2h + 2fJ _


r
(3~,
r2 fJ = fM. (10.14.5)

We obtain then that

2fJ (3~
2h+ - - -dr. (10.14.6)
r r2
The lower limit is not fixed since W is determined up to an arbitrary
additive constant. Using eqs. (13.17) and (13.16) we obtain the equation
for the trajectory and the time

J
r (10.14.7)
~: =± --;::==d=r=== = t - to·
2h + 2fJ _ (3~
r r2

The nature of the sign (plus or minus) depends upon whether r increases
or decreases within the time interval t - to under consideration. In what
follows the sign in front of the radical is omitted and the sign is determined
during the investigation of the obtained relationships.
568 10. Canonical equations and Jacobi's theorem

10.14.2 Keplerian motion in spherical coordinates


Recalling the definition of the Hamiltonian function given in Subsection
10.4.1 and assuming n = 1 we have

1
'2
[(8W)2
8R + R21 (8W)2
8{) +R 2
1
sin 2 {}
(8W)2]
8)", - R h.
f1
=
(10.14.8)

Looking for a solution in the form of eq. (2)

W = (3).,)... + WI (R) + W2 ({}),


we arrive at the equality

R2 [_ (dWI)2 R = (dW2)2
dR + 2h + 2fL] d{) + ~
sin 2 {} ,

which can hold only if both sides are constants. Denoting this constant by
{3~ we obtain two equations

(dWI ) 2 = 2h + 2fL _ {3~


dR R R2'
and the complete integral of the Jacobi-Hamilton equation is given by

(10.14.9)

From the latter we obtain the equation for the trajectory

(10.14.10)

and the equation for time

(10.14.11)
10.14 Separability of variables in the Jacobi-Hamilton equation 569

10.14.3 Motion of a particle in the field of two attracting


centres
This problem has been considered in Sec. 7.7. Let us apply Jacobi's method.
The attracting centres 0 1 and O2 are taken to lie on axis Oz of the cylin-
drical coordinate system r, rp, z the distance 0 1 0 2 being denoted by 2c. Let
the position of the moving particle M in the meridional plane rp = const
be described by the generalised coordinates q1, q2, which are related to the
Cartesian coordinates r, z in this plane by means of the equalities

z = CQ1Q2, r = cJqr -lJ1- q§,

where 1 ::; q1 ::; 00, -1 ::; q2 ::; 1. The curves Q1 = const, q2 = const are the
confocal ellipses and hyperbolas

having the foci at the centres of attraction z = ±c. The sum r1 + r2 and the
difference r1 - r2 of the distances from the attracting centres depend only
on q1 and q2, respectively, since the sum is constant for the ellipse and the
difference is constant for the hyperbola. Thus, these distances are related
to each other by simple relationships. Indeed,
2
r1,2 = (z ± c )2 + r 2 = c2 ( Q1 ± Q2 )2 .
or

The expression for the potential energy has the form

II = _II - 12 = - 1 [(II + h) q1 + (12 - II) q2].


r1 r2 c(Qr-q§)
It is easy to calculate the kinetic energy

T =

The Hamiltonian function becomes

H =
570 10. Canonical equations and Jacobi's theorem

Looking for the complete integral of the Hamilton-Jacobi equations in the


form

we arrive at the equation

8W* ) 2
[ (qi-1) ( - -
(8W* ) 2
+(l-qD - -
{32
+~+~-
{32 1
8ql 8q2 ql - 1 1 - q2

2c [(JI + h) ql + (12 - JI) q2] = 2hc2 (qi - qD '


from which one can conclude that the variables ql and q2 are separable.
Indeed, assuming

yields the relationship

dW
(ql2 -1) ( - d ) 2 + -2
l {3~- - - 2c(JI + 12)ql - 2hc 2ql2
ql ql - 1

- [(1 - qD (ddW2 ) 2 +
q2
{3~
2 -
1 - ql
2c (12 - JI) q2 + 2hc2q~l.
Equating each side to constant -{3, we obtain

(qi-1) (~:lr -{3- qr~l +2c(JI+12)ql+2hc2qi,


(l-q~) (8W2)2
8q2
{3- {3~ +2c(12-JI)q2-2hc2q~.
2
1 - ql
Expression W is now represented in the form

W = (3<pif + J fDf::\
V Fdql) qi _ 1
dql
+
J fDf::\ dq2
V F2 (q2) q~ _ l' (10.14.13)

where
Fdqd = (qi-1) [2hc2qi+2c(JI+12)ql-{3]-{3~,
Fdq2) = (1 - qD [-2hc2q~ + 2c (12 - JI) q2 + {3] - {3~.
The equations for the trajectory and the time take the form

if - {3
<p
J dql
(qi - 1) JFl (qI)
- {3
<p
J dq2
(1 - q~) JF2 (q2)
= a<p
,
-J dql
JFdqd
J
+ dq2
JF2 (q2)
= 2a
,
c2 J qidql - c2
JFl (ql)
J q~dq2
JF2 (q2)
= t - to.

(10.14.14)
10.14 Separability of variables in the Jacobi-Hamilton equation 571

10.14.4 Stiickel's theorem


Stiickel's theorem provides us with the necessary and sufficient conditions
for the separability of variables, that is, for the representation of the integral
of motion (1) in the form (2), [19].
For the sake of notational brevity, we denote

8W
Pk = 8qk· (10.14.15)

Assume that

denotes the complete integral of eq. (1). Then, inserting this expression into
eq. (1) we arrive at the identity which can be differentiated with respect to
each variables (31, ... , (3n-l, h. This yields the following system of equations

(k~l, ... ,n-l), }


(10.14.16)

which is linear for As. By virtue of eq. (13.15) the determinant

8Pl 8Pn 82W


P1 8 (31 Pn 8(31 8(31 8 ql
~= = PI·· ·Pn
8Pl 8Pn 8 2W 82 W
PI 8h Pn 8h 8h8ql 8h8qn
(10.14.17)

is not zero. This system gives

A s -- ~ns
~ (
s = 1, ... ,n ) , (10.14.18)

where ~ns denotes the algebraic adjunct of the element of the n - th row
and the s - th column in the determinant (17). Inserting these values of As
into eq. (1) we find the expression for II

II = h _ ~ ~ ~nk p~ (10.14.19)
2L ~ .
k=1

Due to the above assumption the complete integral of eq. (1) should have
the form (2), hence the k - th column of the determinant (17) can contain
572 10. Canonical equations and Jacobi's theorem

only variables qk, i.e. the following equalities

Pk of3
OPks = 'Psk (qk) (8 = 1, ... ,n - 1) , }
(10.14.20)
OPk
Pk oh = 'Pndqk) (k = 1, ... ,n)

must hold. The n 2 function 'Psk (qk) are arbitrary; the only requirement
is that the determinant obtained by substituting (20) into (17) does not
vanish, i.e.

~= ~o. (10.14.21)

The consequence of eqs. (2) and (15) is the formulae

P~ = (dW)2
dqk
(10.14.22)

On the other hand, equalities (20) allows us to write


n-1
P~ = 2 L f3 s'Psk (qk) + 2h'Pnk (qk) + 2'ljJk (qk) , (10.14.23)
s=l

where 'ljJk (qk) are n arbitrary functions. We obtain

n-1
2 L f3 s'Psk (qk) + 2h'Pnk (qk) + 2'ljJk (qk)dqk.
s=l
(10.14.24)
It is the complete integral since it contains n constants and has a non-
trivial determinant (13.15). In accordance with eq. (17) it differs from the
determinant (21) only by a non-zero factor. Notice also, that the expression
for II, due to eqs. (19) and (23), can be cast in the form

The determinant property

~ ~nk ()
L...J ---;s:-'Psk qk = tins
k=l
yields the following equality

(10.14.25)
10.14 Separability of variables in the Jacobi-Hamilton equation 573

We come to Stackel's theorem: taking n (n + 1) functions i.{Jsk (qk) and


'¢k (qk) we can construct the Hamilton-Jacobi equation in form (1) in which
the coefficients As and the potential energy II are given by formulae (18)
and (25), ~ns being the algebraic adjuncts ofthe elements of the last row of
determinant (21). Formula (24) gives the complete integral ofthis equation.
Conversely, if the complete integral of the Hamilton-Jacobi equation (1) is
represented by the sum (2), then there exist n (n + 1) functions i.{Jsk (qk)
and '¢k (qk) such that the coefficients As and the potential energy II can be
expressed by eq. (1) in terms of these functions and the complete integral
of eq. (1) is given by expression (24).
What remains to be mentioned is that the system of equations yielding
the complete integral of the canonical system of equations, for the case of
the separable variables, has the form

Stackel's theorem fully defines the class of the Hamilton-Jacobi equations


of type (1) with the separable coefficients. However, for a concrete equation
one can give an answer whether this equation belongs to this class only in
simple cases, since difficulties arise when one tries to reconstruct functions
i.{Jsk (qk) and '¢k (qk) in terms of given As and II. According to eqs. (16) and
(20) the question is to find such system of coefficients i.{Jsk that the system

....... ~l.~~~ .(~~~ .~. ~.2.~~~ .(~~~ .~ ....... ~ ~~:'l.~ ~~~~ .~.~: }
A1i.{Jn-l,1 (qI) + A 2i.{Jn-l,2 (q2) + ... + Ani.{Jn-l,n (qn) = 0,
A1i.{Jnl (ql) + A 2i.{Jn2 (q2) + ... + Ani.{Jnn (qn) = 1
(10.14.27)

is fulfilled for given As. This problem can be solved only iffunctions As have
a certain structure. We can reduce the number of appropriate solutions by
requiring that the coefficients of each column depend only on the variable
corresponding to its number.
574 10. Canonical equations and Jacobi's theorem

10.14.5 Liouville's system


As an example leL us consider the Liouville system, that is the system whose
kinetic and potential energies are given by
1
I>s
n
T = 2b (qs) q;, (10.14.28)
s=l

where c and b are the sums of the form


n n
(10.14.29)
k=l k=l

Let us notice in passing that the definition of Liouville's systems given by


eq. (7.7.13) has a less general character. The Hamiltonian function takes
the form

H=~~P; ~ (10.14.30)
2b 6 a s +b'
s=l

and the matrix of functions tpsk satisfying the system of equations (27)

tpn + ... + tp1n 0,


a1 an
tp21 + ... + tp2n 0,
a1 an

tpn1 + ... + tpnn


a1 an

can be chosen as follows


0 0
0 0

a n -1 -an
a n - 1bn - 1 anbn

The row of functions 7/Jk (qk), obtained by means of condition (25), is

7/Jk = -akck·

In formula (24) we have now

(k = 1),
(k = 2, ... ,n - 1),
(k = n).
To simplify the notation we put

{3~ = {31, {3; = {32 - {31'··· ,{3~-1 = {3n-1 - {3n-2, {3~ = -{3n-1,
10.14 Separability of variables in the Jacobi-Hamilton equation 575

so that
n-1
L (3r'Prk (qk) = (3~ak (k = 1, ... ,n) ,
r=1

under the condition


n
L(3~ = o. (10.14.31)
k=1

The complete integral of the Jacobi-Hamilton equation is written as fol-


lows

w= t JV
k=1
2(3k a k + 2hakbk - 2akckdqk, (10.14.32)

where the prime on the coefficients (3k is omitted.


Let us return to the motion of a particle in the field of attraction of two
centres. In the case of plane motion this problem belongs to the Liouville
class with

We obtain then that

where (31 + (32 = O. This equation differs from the above result only in the
notation.
In the general case of a non-planar motion the problem does not belong
to the Liouville class. The system of equations (27) can be cast as

'Pu (q~ - 1) + 'P12 ( 1 - q~) + 'P13 (q~ ~ 1 + 1 ~ q~ ) 0,

'P21 (q~ -1) +'P22 (l-q~) +'P23 (q~ ~ 1 + 1 ~q~) 0,

'P31 (q~ -1) +'P32 (l-q~) +'P33 (q~ ~ 1 + 1 ~q~)


576 10. Canonical equations and Jacobi's theorem

The functions i.ps3 can not depend on the third coordinate i.p, hence they
are constant. Letting i.p31 = i.p32 = -1, i.p33 = 0 we have

and we can satisfy all these relationships by assuming that each of the
values in brackets is constant and that these constants have opposite signs.
For s = 1 and s = 2 we take respectively

i.p12 (1 - q~) - - 112


-q
= 0,
1 2
i.p22 (1 - q~) - --2 = -1.
1- q2

The matrix of functions i.p sk (qk) is given by

1 1
-1
(qr - 1)2 (1 _ q~)2
qr q~
-1
(qr - 1)2 (1 - q§)2
qrc2 qrc2
qr -1 -1- q~
0

Taking, as above, the following values for 'ljJ k

we obtain, due to eq. (24), the expression for W

A change in notation allows us to reduce this expression to eq. (13).

10.15 Keplerian motion


Let us return to the problem of motion of a particle in the field of the
central force of attraction. We introduce the system of polar coordinates
10.15 Keplerian motion 577

r, <p which lies in the plane of the trajectory and has origin at the centre of
attraction. Then the solution is given by formulae (14.7)
.,.

± JJ dr
f (r) = t - to, (10.15.1)

where

f(r) = 2h+ 2p, _ (3~. (10.15.2)


r r2
The momenta are determined in terms of the coordinates by differentiating
eq. (14.6)

(10.15.3)
Function f (r)
under the square root must be positive. Hence, equation
r 2f(r) 0 should have only real-valued roots rl,r2, otherwise r 2f(r),
=
being negative at sufficiently small r, would remain negative for all real r.
Assuming rl < r2 we can write the following
(10.15.4)
where

(10.15.5)

The signs of the roots are different for h > 0 whereas both roots are positive
for h < O. We restrict our attention to the case when h < o. As follows
from eq. (4)
(10.15.6)
The generic names for the points of the orbit around any attracting
centre corresponding to the distances rl and r2 from the attracting centre
are pericentron and apocentron. Particularly, these points are referred to as
the perihelion and aphelion for the orbits around the sun and the perigee
and apogee around the earth.
Let at t = 0
p.,. = p~ = TO < O.
Then the distance r decreases within a certain time interval (0, to) from
the initial value ro to rl. Thus a minus sign should be taken in the first
equation in (1) and we obtain
.,.
J
dr
- v7T.T5 = t and (10.15.7)
"'0
578 10. Canonical equations and Jacobi's theorem

where to denotes the time when the pericentron is reached. The value of to
is finite, as the denominator of the integrand is proportional to y'r - rl.
Counting t from the time instant to we should take rl as the lower inte-
gration limit and the plus sign in front of the integral. Denoting the time
when the apocentron is reached by tl we have

JVJTT5
T
dr
= t - to and
Tl

Beginning with the moment h we again have

-J~=t-h
T

and
Tl

where t2 stands for the new time instant when the apocentron is reached.
As shown below the return to the pericentron and apocentron corresponds
to the same points of the orbit and the increment in 'P is 27r. This ensures
that the motion is periodic and has a closed orbit. The time interval

(10.15.8)

is equal to the half-period of the orbital motion.


Estimation of the integrals in formulae (1) is based on the introduction
of a new variables w instead of r by means of the relationship

r = rl cos
2W
"2 + r2 sm "2 = '12 (rl + r2) - '12 (r2
.2w
- rd cosw. (10.15.9)

It is easy to prove that inequalities (6) are satisfied for any real w. The sub-
stitution (9) is successfully applied in any cases when it is known in advance
that the quantity in question lies between two limiting values (rl and r2 in
our case). In Kepler's problem w is referred to as the eccentric anomaly. In
the pericentron and apocentron it takes values 0 and 7r, respectively.
Equality (9) can be written in a simpler way provided that the following
notation

(10.15.10)

is introduced, where 0 < e < 1. Then we obtain

r = a (1 - e cos w) . (10.15.11)

Equalities (5) and (10) enable considerable simplification of eq. (4), namely

r2 f (r) = {Jae 2 sin 2 w,


10.15 Keplerian motion 579

and relationships (1) reduce to the form

j
T rdr
r:::?:TT::\ =
ava jW (1 - e cos w) dw = t - to,
r,; (10.15.12)
V 1'2 f (1') V f.L
Tl 0

T W

CP=Jf.L a (1-e 2 ) j dr =~j dw . (10.15.13)


rJ17i) 1- ecosw
Tl 0

Here

(10.15.14)

and Ct<p = 0, which implies that cp is measured from the direction from the
attracting centre to the pericentron. Angle cp is called the true anomaly.
By virtue of eqs. (12) and (8), we find

2n ,jjj
n-----
- T - ava' (10.15.15)

These equations for the period T and frequency n, termed the mean motion,
expresses Kepler's third law. Returning to eq. (12) we come to Kepler's
equation

w - e sin w = n (t - to) = (, (10.15.16)

which serves to determine the eccentric anomaly as a function of time. The


value of ( is called the mean anomaly.
Integrating eq. (13) we obtain the formula which relates the true anomaly
cp to the eccentric anomaly w

cp = 2 arctan ~
- - tan -,
+e w cp =
tan- ~+e w
--tan-. (10.15.17)
1-e 2 2 1- e 2

From this formula we find


e + cos cp sin w = ~...,----_sl_·n-,cp,---_
cosw = , (10.15.18)
1 + e coscp 1 + ecos cp

and the inverse formulae


cosw - e ~ sinw
cos cp = , sin cp = V 1 - e 2 . (10.15.19)
1- ecosw 1- ecosw
The orbit is seen to be an ellipse with one focal point at the attracting
centre, the constants a and e being the semi-major axis and the eccentricity
580 10. Canonical equations and Jacobi's theorem

//,.,

.r,.,

FIGURE 10.2.

of the ellipse, respectively. Indeed, replacing cos w in eq. (11) due to eq.
(18), we arrive at the equation for the ellipse in polar coordinates

a (1 - e2 ) p
r = -'-----"- (10.15.20)
1 + e cos<p 1 + e cos <p

The origin of the polar coordinate system is at the focal point F and p
denotes the parameter of the ellipse, see Fig. 10.2.
The obtained solution, given by eqs. (20), (16), (17), contains three con-
stants a, e, to. However, the general solution of the motion of the particle
in space should have six constants. Let us take three Euler's angles as the
remaining constants. Euler's angles determine the position of the orbital
plane and the axes in this plane relative to the axes O~TJ( fixed in the space
and having their origin at the attracting centre.
Let us determine the directions of axes Oxyz by the unit vectors el, e2, e3.
The vector el is directed from the attracting centre to the pericentron, e2
lies in the orbital plane perpendicular to el and points to the direction of
<p , and e3 = el x e2 is orthogonal to the orbital plane. The Euler's angles
are denoted by D, i, w, see Fig. 2.2. Angle D determines the direction of
the straight line which is the intersection of plane O~TJ and the trajectory
plane. This line is the nodal line, point N and angle D are referred to as
the ascending node and the longitude of the ascending node. Angle i de-
scribes the angle between the orbital plane and the plane O~Tf. Angle w is
the angle between the nodal line (pointing to the ascending node) and axis
Ox (pointing to the pericentron).
10.15 Keplerian motion 581

Putting nOW the radius vector of the moving particle in the form

r = elr cos <p + e2rsin<p = a [e 1 (cosw - e) + e2~ sinw]


(10.15.21)

and projecting it onto the axes of the system O~'Tf( we arrive at expressions
for the coordinates ~,'Tf, ( of the particle which contain the following six
constants

a, e, to, 0, i, w (10.15.22)

called the elliptic orbital elements. These expressions can also be obtained
in spherical coordinates if we usc the solutions (14.10) and (14.11), [44].
Recalling that, due to eqs. (16), (15) and (11),

. n (iiI
w = 1- ecosw = V~-:;:' (10.15.23)

we find, by differentiating expression (21), that the velocity vector of the


particle

v = ~ ( -el sin w+ e2 ~ cos w) . (10.15.24)

The value of the velocity vector and the angle with the radius vector rare
given by the equalities

v=/¥ 1 + ecosw
l-ecosw'
cos Q
r·v
= -- =
rv vI -
esinw
---r===;;=::::;;:::=
e2 cos 2 w
(10.15.25)

Now it is easy to obtain the constants a, e, Wo in terms of the initial values


ro, Vo, Q. We have
2{3 -1
a ro
= 2 (1- (3)'
e = . fcos 2 Q
V
+ (2{3 - 1)2 sin 2 Q, cOS W o = ---.
e
(10.15.26)

Here
2
(3= rovo. (10.15.27)
2J-t

If the attracting centre is the earth then, by replacing J-t = f Min eq. (5.6.7)
by its value gm according to eq. (5.6.7) and denoting

Voo = J2gRo, (10.15.28)


582 10. Canonical equations and Jacobi's theorem

Ro being the radius of the earth, we obtain

{3 =!2.- (~)2 (10.15.29)


Ro Voo

The value Voo is referred to as the escape velocity, i.e. it is the velocity
needed by a particle to reach infinity. If a particle possesses a horizon-
tal velocity (cosa = 0) equal to JgRo = 0.707v oo then, as follows from
formulae (26), the orbit is a circle with the radius of the earth.
Depending upon the initial velocity Vo the elliptic motion beginning on
the surface of the earth (TO = Ro) can be divided into two types. The" earth
motions", for which

0::; vo::; JgRo, 0::; {3::;~, ~Ro::; a::; Ro, }


1 ~ e ~ cosa, 7r ~ Wo ~ i, (10.15.30)

belong to the first type. These motions take place along the ellipses inter-
secting the surface of the earth at the start and fall points. The start point
is close to the apogee, that is the centre of the earth lies in the far focus
of the orbit and the second focus is close to the surface of the earth. The
parabolic orbit in the homogeneous field of the earth's gravity (Ro = (0)
corresponds to the left limiting case of inequality, i.e. {3 = 0, e = 1, Wo = 7r.
The elliptic orbits of the second type are described by the inequalities

JgRo::; Vo ::; v oo , ~::; {37r::; 1, Ro::; a::; 00, }


(10.15.31)
cos a ::; e ::; 1, "2 ~ Wo ~ O.

The focus which is the centre of the earth is closer to the perigee than
the second focus which is removed from the start point on the distance
greater than the earth radius. Such orbits are typical for earth satellites.
The left limiting case of inequalities (31) describes the above circular orbit
in the case of the horizontal start (cos a = 0), whilst the right limiting case
corresponds to the parabolic orbits with the focal point at the centre of the
earth. If Vo > v oo , i.e. {3 > 0, the orbit is no longer elliptic and the orbit
becomes a hyperbola.
Let us consider Kepler's equation. Among many possible solutions, the
expansion of the difference between the eccentric and mean anomalies

w-(=esinw (10.15.32)

using trigonometric series

L
00

w- ( = ak sin k( (10.15.33)
k=l
10.15 Keplerian motion 583

is proved to be more efficient.


This representation is possible since the increment of 27f in ( leads to
the same increment in w. In addition to this, Kepler's equation does not
change when both ( and w change sign. Thus, w - ( is an odd periodic
function of ( which can be represented by series (33). The coefficients are
as follows

J
7r

ak = ~ (w - () sink(d(.
o
Integrating by parts and using Kepler's equation we obtain

2
7fk J
7r

o
2
cosk(dw= tfk J
7r

0
.
cosk(w-esmw)dw.

The functions determined by the integrals of the type

In 1
(x) = :; J
7r

cos (nw - xsm w) dw . (10.15.34)


o
were first introduced by Bessel for solving Kepler's equation. Hence,

_ r 2~ Jk (ke) . k r
w-.,+ ~ k sm.,. (10.15.35)
k=l

Kepler's equation yields

. _2~Jdke). k r (10.15.36)
smw - - ~ k sm.,.
e k=l
This way of calculation is applicable to other functions of w. For instance,
cos w is the even periodic function of ( and can be cast in the form
b <Xl

cosw = ; + Lbkcosk(.
k=l

The coefficients are given by

~ = :;
2" IJ 7r

coswd( IJ
=:;
7r

cosw (1 - ecosw) dw = -"21 e,


o 0
584 10. Canonical equations and Jacobi's theorem

~ ~

~=~J~w~~~=!J~~~w~

[1 1
o 0

~ :k c'" (k( - w) dw - cos(k( + w) dw 1


~ :k {1 cos I(k -1)w - keffinwl dw - 1 cosl(k+ 1) w - kee'nwl dW}'

Hence, by means of eq. (34)

1 1
cosw = -"2e + L
00

k [Jk-l (ke) - Jk+l (ke)] cos k(. (10.15.37)


k=l
11
Perturbation theory

11.1 Method of parameter variation


Along with the given system of equations of motion
. aH . aH + Qs
qs = -a
Ps
' Ps = --a
qs
(8 = 1, ... ,n) (11.1.1)

we consider an auxiliary (simplified) canonical system


. aHa . aHa
qs = aps ' Ps = --- (8=1, ... ,n). (11.1.2)
aqs
It is assumed that the Hamiltonian function Ha of the system of equations
(2) is chosen in such a way that we can find a general solution containing
2n constants
qs=qs(t;O!l, ... ,O!n,/31'··· ,/3n ), } (11.1.3)
Ps = Ps (t;O!l, ... , O!n,/31'··· ,/3n ) (8 = 1, ... ,n).
By analogy with Sec. 10.12 we assume that this system of equations is
solvable for O!k, 13k

O!k =O!k(ql, ... ,qn,Pl,··· ,Pn;t), }


(11.1.4)
13k = /3dql, ... , qn,Pl,··· ,Pn; t) (k = 1, ... , n) .
Then, as pointed out in Sec. 10.12, any of the Poisson brackets

(11.1.5)
586 11. Perturbation theory

is either constant (equal to zero, in particular) or expressed in terms of


the values D:k, (Jk' none of the brackets depending explicitly upon t and
variables qs,Ps. By virtue of eq. (10.6.3) we can write down 2n identities

(11.1.6)

Let us proceed now to the given system of equations (1). The idea of the
method of parameter variation is that the general solution of the original
equations (1) is sought in the same form (3) however it is assumed that
D:k, (Jk are no longer constants but functions oftime. Then equations (3) can
be deemed as the formulae transforming the new variables D:i, (J i into the
old variables qs,ps, whilst equations (4) are understood as the formulae for
the inverse transformation. The concern here is to construct the system of
differential equations for the new variables provided that the old variables
are solutions of the initial equations (1). If the old variables were solutions
of equations (2), then the "new" variables would be constant.
The identities (6) are valid since they are derived from equations (4)
which do not change. For the same reason the Poisson brackets (5) do not
change their form either. .
We need expressions for the time-derivatives of the new variables with the
system of differential equations (1) being taken into account. Differentiating
expressions (4) yields

Here the identities (6) are taken into account and qk, Pk are replaced by eq.
(1). By analogy, we obtain ils and arrive at the system of equations

(11.1. 7)

referred to as the equations of perturbed motion. Clearly, their right hand


sides are assumed to be expressed in terms of D:k, (Jk by means of formulae
(3).
We can cast eq. (7) in another form by using formula (10.7.14) for the
transformation of the Poisson brackets. In this formula we should replace
11.1 Method of parameter variation 587

Qs, Ps and u, v by as, (3s and as, H - Ho. Then we have

Taking into account that

we have the equality

By analogy

The system of differential equations of perturbed motion (7) is now ex-


pressed in the form

~ [8(H - Ho)
.
as = ~
k=l
8
ak
(as, ak) +
8(H - Ho)
8{3
k
(as, (3k) + a
8as ]
Pk
Qk ,

(3. = ~ [8(H - Ho) ({3 ) 8(H - Ho) ({3 (3) 8{3sQ]


s ~ 8 s' ak + 8{3 s' k + 8 k
k=l ak k Pk
(8 = 1, ... ,n).
(11.1.10)

The special convenience of this notation is that the Poisson brackets are
either constant or expressed in terms of ak, {3k.
The analysis up to this point has been based upon the replacement of the
system of differential equations (1) by another system (10). On assuming
the solution of the latter system to be known we immediately obtain the
solution of the initial problem (1). It is evident that this approach can be
fruitful when the system (10) is more easily solved (at least approximately)
588 11. Perturbation theory

than the initial one. An approximate solution will be effective if the aux-
iliary system of equations (2) can be chosen in such a way that it differs
from the initial system (1) only by small secondary terms. Then quanti-
ties as, f3s, which are constant in the solution of the auxiliary system of
equations (2), will differ slightly from the constants in problem (1). This
is immediately seen from equations (10) since the time-derivatives as, /3 s
have the same order as the presumably small H -Ho, Qk. Thus, as, f3s turn
out to be slowly changing functions of time and the numerous approximate
methods, in particular the method of successive approximations, become
applicable for integrating the system (10) of differential equations of per-
turbed motion. Denoting for brevity the right hand sides of eq. (10) as

J
t

<I>s (alf3;t) and 'lis (alf3;t), we can replace ak,f3 k in the relationships

°
J J
t t

as = a~ + <I>s (alf3; t) dt, f3 s = f3~ + 'II s (alf3; t) dt, (11.1.11)

° °
(obtained directly from (10)) by their initial values aZ,f3Z at t = O. This
replacement of the slowly varying quantities by constants cannot lead to
considerable discrepancy, at least if t is not large. The problem is reduced
to quadrature

J J
t t

as = a~ + <I>s (aolf3°; t) dt, f3 s = f3~ + 'lis (aolf3°; t) dt. (11.1.12)

° °
The process can be continued, that is the values as, f3 s determined from eq.
(12) can again be substituted into eq. (11) and so on. We do not touch upon
the problem of convergence. In a number of cases this approach remains
valid even if the process does not converge. Evidence for this is provided
by the practice of astronomical calculations, see [24].
When the first approximation is sought, one uses approaches based upon
the replacement of the right hand sides of the equations for the perturbed
motion (10) by their average values over a certain time interval T

= TJ
T
-
<I>s(alf3) 1 <I>s(alf3;t)dt,

°
with ak, f3k on the right hand sides being assumed constant while the inte-
grals are estimated. The problem reduces to consideration of the following
system of equations

as = <I>s (alf3), /3s = q,s (alf3) (8 = 1, ... , n) (11.1.14)


with small right hand sides which are independent of time.
11.2 Canonical equations of perturbed motion 589

11.2 Canonical equations of perturbed motion


The system of equations for the perturbed motion is simplified in the case
when the solution of the auxiliary system of equations (1.2) represents a
canonical transformation of quantities ak, (3 kinta qs, Ps. As mentioned in
Chapter 10, this takes place in two cases: firstly, when eq. (1.3) is the
solution of Cauchy's problem for the system of differential equations (1.2),
i.e. as, (3s are the initial values of the variables qs,Ps and, secondly, when
solution (1.3) is the general integral of the canonical system (1.2) obtained
from the complete integral of the Jacobi-Hamilton equation.
According to eq. (10.7.12) for the canonical transformation

(11.2.1)

which results in a drastic simplification of the form of equation (1.10) for


the perturbed motion. They take the form

. _ o(H - Ho) ~ oa sQ
as - 0(3 +~ 0 k (s=l, ... ,n), )
s Pk
k=l
(11.2.2)
(3. --
- o(H -Ho) ~ o(3sQ
+~- k (s=l, ... ,n).
s oas k=1 0Pk

In particular, if

Qk = 0 (k = 1, ... , n) ,

i.e. when equations (1.1) are canonical, then the equations for the perturbed
motion

. 0 (H - Ho) (3. s = 0 (H - Ho)


(s = 1, ... , n) (11.2.3)
as = o(3s' oas

are also canonical.


Replacing Qs, Ps in relationships (10.8.16) by as, (3s, respectively, we can
write eq. (2) as follows

(s ~ 1, ... , n). )
(11.2.4)

In order to write down equations (4) there is no need to know the inverse
transformation (1.3) and this is their advantage over equations (2).
590 11. Perturbation theory

If the generalised forces Qs depend only on time (in particular, constant)


then the equations for perturbed motion take the canonical form

(8 = 1, ... ,n). (11.2.5)

Of course, this can be easily foreseen as the initial equations of motion


(1.1) are cast in the form of the canonical equations with the Hamiltonian
function
n
H* = H - LQkqk, (11.2.6)
k=l
see Sec. 5.3. for the definition of the generalised potential energy.
The method of parameter variation was suggested by Lagrange in [55].
The equations of perturbed motion are derived in form (1.10) as well as in
form (3), however Lagrange was not familiar with the canonical equations.
He characterised the essence of the method with the following words: "in
mechanical problems, the first solution is usually sought by taking into con-
sideration only the main forces acting on the bodies. In order to spread the
solution on other forces which can be termed perturbing it is easy to pre-
serve the form of the first solution, but to consider the arbitrary constants
as variable quantities. If the values which have been neglected earlier and
should be taken into account now are small, then the new variables will
be nearly constant and the conventional methods of approximation can be
applied to solving the problem".
The outstanding mathematicians of the first half of the nineteenth cen-
tury developed these ideas by Lagrange.

11.3 Motion of a particle in the gravitational field


of the rotating Earth
The position of the particle is described with respect to the system of the
"start" axes Oxyz, whose origin is placed at point 0 and is determined on
the surface of the earth by the polar angle {)o (angle complementary to the
latitude) and longitude >'0. Axis Oz is directed from the earth centre () to
point 0 along the radius vector Ro.Axes Ox and Oy lie in the tangent
plane to the sphere of radius RQ at point 0, the first axis is directed along
the tangent to the meridian of the sphere pointing to the side of decreasing
angle {)o whilst the second one is directed along the parallel circle pointing
to the side of decreasing angle >'0. In other words, these are the northern
11.3 Motion of a particle in the gravitational field of the rotating Earth 591

and western directions provided that the non-sphericity of the earth is


neglected.
We begin by determining the Hamiltonian function K, which is given
by expression (10.11.9) and then constructing the canonical equations of
motion (10.11.10). The expression for the kinetic energy of the particle of
unit mass is required to be written in the form (11.10.4). Noticing that in
our case

where, as always, U denotes the angular velocity of the earth we have

w· (r' x ~~) = U(ilsin'!9o +i3cos'!90)' (r' XiI) = -Uy cos '!90,

and by analogy

w. (r' x~~) = U (x cos '!90 - z sin '!90), w . (r' x~~) = U y sin '!90.

Then by virtue of eq. (10.11.4) we obtain

T; = ~ {(PI + Uy cos '!9 0)2 + [P2 - U (xcos'!9 o - zsin'!90)]2 +

(P3 - Uy sin '!90)2} . (11.3.1)

Let us proceed to calculate the components of expression (10.11.9) for


the Hamiltonian function K. For m = 1 we have
w . e . w = r P w2 - (r' . w) 2 = U 2 (x 2 + y2 + z2) - U 2 (x sin '!9 0 + z cos '!90) 2
= U 2 [(z sin '!90 - x cos '!9 0)2 + y2] ,

(w x vo) . r' U [(il sin '!90 + i3 cos '!90) x (-i2R~U sin '!9 0)] . r'
U 2 R~ sin '!9 0 ( - z sin '!9 0 + x cos '!9 0) ,
as the velocity vector Vo of the pole of the start system of axes equals

* can be cancelled out and we


As it is a constant value, the term with Vo
obtain

- ~w. e· w + mew x vo)· r' = _~u2 [(zsin'!9o - xcos'!90)2 +


2 2
2R~ (z sin '!9 0 - x cos '!9 0) + R~2 sin 2 '!9 0 + y2] +
~R*2U2sin2'!9
2 0 0
=-~U2h2+const
2 '
592 11. Perturbation theory

where the constant term can be omitted and h 2 denotes the square of the
distance from the point to the axis of rotation of the earth

According to eq. (5.6.2) the value

1 1
11* = II - "2w. e· w + m (w x vo) . r' = II - "2U2h2 + const, (11.3.2)

is the potential energy of the weight, that is the sum of the potential ener-
gies of the gravity force and the centrifugal force. 11* is given by eq. (5.6.9)
and is expressed in terms of the distance r from the centre of the earth and
angle {J which is complementary to the latitude of the actual position of
the particle. We have

r = ro + r, • Tl*'
= 13 .... "0 + IIX + 12Y

+ 13Z,

cos {J = -.
r
r (. . {J
II sm 0 + 13• cos {J 0 )
and thus

r2 = X2 + y2 + (R(; + z)2 , cos{Jo = ~ [x sin {Jo + (KO + z) cos {Jol.


r

Here, in accordance with eq. (5.6.11)

R(; = Ro - ~aRoP2 (cos{Jo) ,


Ro and m
denoting the mean radius of the earth and the radius of the
normal spheroid at point 0, respectively. Then

r
Ro

x y z
Here Ro' Ro' Ro are assumed to be small values of first order, whereas
a (and correspondingly m) are values of second order. Terms up to third
11.3 Motion of a particle in the gravitational field of the rotating Earth 593

order are kept. Then we have

While calculating the terms in eq. (5.6.9) mUltiplied by m or a it is sufficient


to retain the terms of first order. Taking into account eqs. (5.6.3) and
(5.6.7), we obtain

where P2 (cos 'l9) should be replaced by the following expression

dP2 (COS'l9))
P 2 (cos'l9) P2 (cos'l9 o) + ( dcos'l9 0 (cos'l9 - cos'l9 o)

x .
P2 (cos 'l9 0 ) + 3 flo cos 'l9 0 sm 'l9 0 ·

Inserting this into eq. (5.6.9) yields

(11.3.3)

where f3 denotes Clairaut's constant determined by eq. (5.6.19).


594 11. Perturbation theory

The Hamiltonian function K is represented in the form

K = ~ (pl + pi + pj + 290 Z ) + {U [(PI cos'!9o - P3 sin'!9o) y-


x 2 + y2 _ 2z2}
P 2 (xcos'!9 o - zsin'!9o)] + 90 2Ro +

{ '2I U 2 [ y 2 + (x cos '!90 - .


zsm'!9o) 2 4
- "3Roz ] -

90 [
3Z (x 2 + y2) - 2z 3
2R~
2
- "3f3ZP2 (cos'!9o) - 20:xcos'!9osin'!90
1}+ ...
= K(O) + K(I) + K(2) + ... (11.3.4)

In what follows, term K(O) determined by compression of the geoid and the
influence of the centrifugal forces cased by rotation of the earth is neglected
because their influence is comparable with the attraction of the particle by
the moon, see [1]. The Coriolis force, being linear in the angular velocity
of the earth, is taken into account by K(I) which is assumed to be of first
order of smallness.
Let us determine the simplified system of equations by the component
K(O), then we have

These are the equations of motion of the particle without the influence of
the rotating earth. Their solutions are given by

PI = f31,
(11.3.6)
X = f3 l t + 0:1,

where O:i and f3i are the initial values of the coordinates and momenta of
the system of equations (5).
The equations (2.3) for perturbed motion are constructed by means of the
Hamiltonian function K(I) in which the coordinates and momenta should
11.3 Motion of a particle in the gravitational field of the rotating Earth 595

be replaced by expressions (6)

K(l) = U [(,61 cos 79 0 - ,63 sin 79 0 ) (,62t + (2) - ,62 (,61 t + ad cos 79 0 +

got (,62t + (2) sin 790 + ,62 (,63t + a3 - ~got2) sin 79 0] +

2~0 [(,61 t + ad 2 + (,62t + (2)2 - 2 (,63 t + a3 - ~got2) 2].


(11.3.7)

We obtain the following system of three canonical equations of perturbed


motion for quantities ai

(11.3.8)

and three equations for ,6i

(11.3.9)

Replacing ai and ,6ion the right hand sides of these equations by their
initial values a? and,6? due to eq. (1.12), we arrive at approximate solutions
of the form
596 11. Perturbation theory

and so on. Inserting these into eq. (6) we obtain the expressions for the
coordinates

(11.3.10)

and momenta

(11.3.11)

The relative generalised momenta which, for a particle with unit mass,
are equal to the projections X, y, z of the velocity on axes Oxyz can be
calculated by means of formulae (10.11.13), to give

x= PI + Uycos1)o ~ PI + U (ag + ,8~t) cos 1)0, )


y = P2 - U (x cos 1)0 - zsin1)o) ~ P2 - U (a~ + ,8~t) cos 1)0+
U (ag + ,8~t) sin 1)0 - ~ U 90t2 sin 1)0,
z = P3 - Uy ~ P 3 - U (ag + ,8~t) sin 1)0.
(11.3.12)

The same relationships can be obtained by differentiating equalities (10).


The constants a~, ag, ag are the initial values of the coordinates whereas
,8~, ,8~,,8~ are determined in terms of the initial values of the velocities by
means of the following relationships
. (./0
xO=fJI+
U a2cos
0 v.0 O, • (./0
y=fJ2-
U aIcos
0 v.0 O+ U a3' (./0 U a2s1nvo·
o ·ZO=fJ3- 0 . .0

For example, let us consider a particle falling near the surface of the
earth from height h without any initial velocity, then

,8~ = 0, ,8~ = -Uhsin1)o, ,8~ = o.


11.3 Motion of a particle in the gravitational field of the rotating Earth 597

Neglecting the terms depending upon the change in the gravity force with
height, we obtain due to eq. (10)

x = _U 2ht2 sin'l'Jo cos 'l'Jo ~ 0, y = - ~ gt 3 sin 'l'J o, }


(11.3.13)
Z = h- ~gt2 + U 2ht2 sin2 'l'Jo ~ h - ~gt2.
Keeping the terms proportional to U 2 in the latter equality would make
no sense under the adopted accuracy of the calculation. As a result we
obtain an expression for the eastern deviation of the falling particle. The
initial conditions for the particle thrown vertically upward are as follows
a~ = 0, ag = 0, ag = 0, {3~ = 0, {3g = 0, {3g = zo,
and, from eq. (10), it follows that
. 1 2
x = 0, y = U zot 2 sin'l'Jo - ~Ugt 3 sin'l'Jo,
2
= zot
(11.3.14)
Z - -gt

At the time instant when the maximum height is achieved, i.e. when z = 0,
the value of the western deviation is
* 2 z5 .
y = "3Ug2 sm'l'Jo. (11.3.15)

Let us consider another case of firing a particle from the surface of the
earth with the initial velocity vo comprising angle 8 with the plane of the
horizon. Then the projection of vo on this plane comprises angle c with the
northern direction and in this case
a~ 0, ag = 0, ag = 0,
{3~ Vo cos 8 cos c, {3g = Vo cos 8 sin c, (3g = Vo sin 8.
Formulae (10) yield
x = vot cos 8 cos c + Uvot2 cos 8 sin c cos 'l'Jo,
Y = vot cos 8 sin c - U vot 2 cos 8 cos c cos 'l'J o+
U vot 2 sin 8 sin 'l'Jo - ~U got 3 sin 'l'J o, (11.3.16)

Z = vot sin 8 - U vot 2 cos 8 sin c sin 'l'Jo - ~got2.


Placing axis O~ in the plane of throwing and axis Ory perpendicular to
it, so that the angle between axis Ory and the western direction is c, we
obtain

~ vot cos 8 + Uvot2 sin8sin'l'Josinc - ~U9ot3sin'l'Josinc,


1
ry -Uvot2 (cos 8 cos 'l'Jo - sin 8 sin 'l'Jo cos c) - "3U got 3 sin 'l'Jo cosc.

The latter equality determines the deviation from the plane of throwing.
598 11. Perturbation theory

11.4 Motion of a particle in a resistive medium


This case was analysed by Hamel in [36]. The influence of the rotating earth
and change in the gravity force with height are neglected. Then directing
axis x horizontally and axis y along the upward vertical we have for the
particle with unit mass

H ="21 (2
Px + P2 )
y + 2gy .

The resistance force acts in a direction opposite to the velocity of the


particle and its projections on axes Ox and Oy are given by

Fx
X
= --f(v), Fy =
iJ
--f(v),
v v
where x= Px, iJ = Py· Neglecting the resisting force we obtain

x = al + (31t,
The equations for the disturbed motion (2.4) are thus as follows

0: 1 = t(31 f (v), 0:2 = t(32 - gt f (v),


v V

/31 = _(31 f (v), /32 = _(32 - gt f (v); (11.4.2)


v v
v = J(3~ + ((32 - gt)2.

Provided that the resistance is small, we can assume ai, ... ,(32 on the
right hand sides of these equations are constant. Then

where

For the sake of integration we introduce the new variable oX

T = ! ((3g - (3~ sinh oX) •


9
11.5 Influence of small perturbations on oscillations about the equilibrium 599

Then
(30
dT = - --1.. cosh >'d>'
9
and the solution of the equations in (1) can be cast in the form

°
X= a~ + (3~t + (~1) /
2 I-'
(sinh>' - sinhp,) f ((3~ cosh >.) d>',
1-'0
Y= ag + (3gt - 1 ((30)2/1-' (sinh>' -
"2gt2 + gl sinhp,) f ((3~ cosh>.) d>',
1-'0
( 11.4.3)

where

For example, the case for a quadratic resisting force f (v) = cv 2 causes no
problem for calculation despite the cumbersome expressions.

11.5 Influence of small perturbations on


oscillations about the equilibrium
In this section we discuss the difficulties associated with the method of
successive approximations and show how to overcome them by applying
the averaging method. For the sake of simplifying the analysis we apply
the approach to a particular case which is the problem of the influence
of small perturbations on oscillations of a system about its equilibrium
position.
Let us consider a material system with stationary constraints which can
exhibit small motions about the equilibrium position. Denoting the gener-
alised coordinates measured from the equilibrium position by qs we repre-
sent the expressions for the kinetic and potential energies in the form

(11.5.1)

where
600 11. Perturbation theory

ask denoting the constant component of Ask. First order terms are absent
in the expression for the potential energy since the equilibrium corresponds
to zero generalised coordinates, thus, due to eq. (6.5.11)

(all) _
aqs 0 -
0
.

The power series for II* begins with terms of order not lower than the third.
It is known that the following two quadratic forms

one of them being positive definite, can be simultaneously transformed to


forms with no products of different variables with the help of a linear trans-
formation, [56]. These new coordinates are called principal coordinates. Let
us assume that the transformation to principal coordinates has been car-
ried out and use the same notation qs for the principal coordinates. Then,
instead of eq. (1), we obtain the expressions

(11.5.2)

The small oscillations are assumed to take place about the position of
a stable equilibrium, then II is a positive definite function, at least for
sufficiently small Iqsl and all the coefficients are also positive. Of course,
the coefficients as are positive, too. Their ratios

"s
,2 = Cs (1
,s = , ... , n ) (11.5.3)
as

denote the squares of the frequencies of the principal oscillations of the


system about the position of the stable equilibrium under consideration.
Along with the potential forces, non-potential forces act on the system.
In general, the generalised forces Qs corresponding to the non-potential
forces depend on the coordinates qs, the generalised velocities qs and time.
The momenta ps are introduced by the relationships

(11.5.4)

This is a system of linear equations from which the generalised velocities


are expressed in terms of the momenta. With this in mind, the generalised
forces Q s are seen to be expressed in terms of the generalised coordinates,
momenta and time.
11.5 Influence of small perturbations on oscillations about the equilibrium 601

It is natural to take the Hamiltonian function of the auxiliary problem


(1.2) in the following form

(11.5.5)

Then the difference H - Ho is determined by the equality

in which <is are assumed to be expressed in terms of the momenta with the
help of the system of equations (4). Omitting the first (ordinary) sum in
this expression would be an error.
The simplified canonical system (1.2) is as follows

.
qs P
= - ,s.
Ps = -csqs (
s =1
, ... )
,n (11.5.7)
as
with Cauchy's integral in the form

qs = Ct s cos Ast + 11 ~ sin Ast, Ps = -CtsasAs sin Ast + 11s cos Ast,
as/\s
(11.5.8)

so that Ct s and 11 s are the initial values of the generalised coordinates and
momenta, respectively. Equations for perturbed motion can be cast in the
form (2.4)

. _ a (H - Ho) _ Q ( I . ) sin Ast }


Ct s - £l(.l s q P, t "
UfJs as/\s
(11.5.9)
. a(H-Ho)
I1s = - aCts + Qs (qlp; t) cos Ast,
where H - Ho and Qs are expressed in terms of Ct s , I1s by means of eq. (8).
This method of integration of this system of equations based upon re-
placing the unknown variables Ctk,l1k on the right hand sides by constant
values is not applicable to the problems of oscillations. It provides us with
a solution which is quantitatively correct for a sufficient small time inter-
val, however the qualitative character of the motion remains obscure. The
method of averaging of the right hand sides of equations (9) is more suit-
able to this end. This method belonging to the more important class of
solution of problems of nonlinear mechanics was suggested by Krylov and
Bogolyubov [54] and developed in numerous papers, see e.g. [12].
With the intention of explaining the reasons behind the apparent diffi-
culties, we limit our consideration to a linear system with two degrees of
602 11. Perturbation theory

freedom. This enables us to omit H - Ho. Under perturbing forces we un-


derstand those resisting forces, which are linear function in the velocities
are determined by the positive definite dissipation function

(11.5.10)

where the coefficients b1 , h, b2 are small. Further analysis is carried out up


to terms linear in these coefficients.
The generalised velocities in the generalised forces

should be expressed in terms of the momenta. Then we obtain

(11.5.11)

The equations of the perturbed motion (2.4) take the form

Substituting the expressions (8) for the momenta into the right hand side
of eq. (12) and replacing in the obtained expressions the products of the
circular functions of the sort cos Al t cos A2 t by the circular functions of the
sum and difference of the arguments, we arrive at the equations

. b1 b1
Ct1 = --Ct1 + -Ct1 cos 2A1t+
2a1 2a1
hA2
\ Ct2 [cos (AI + A2) t - cos (AI - A2) tJ +
-2
aU\l

~\ /31 sin2A1t + 2a1 ha 2 A1 /32 [sin (AI + A2) t + sin (AI - A2) tJ,
2alAI
. b1 b1 1.
/3 1 = --/31 - -/31 COs2A1t + -b1A1Ct1 sm2A1t+
2a1 2a1 2
h~2 Ct2 [sin (AI + A2) t - sin (AI - A2) tJ -
h
-2 /32 [cos (AI + A2) t + cos (AI - A2) tJ .
a2
(11.5.13)

By analogy we obtain equations for 002, fJ2.


11.5 Influence of small perturbations on oscillations about the equilibrium 603

If we replace CY s , f3 s on the right hand sides of these equations by constant


values cy~, f3? the result of the integrations contains the terms

which are referred to as the secular terms. Such solutions appropriate for
short intervals of time represent the first terms of the expansion of the
exact solution of the series in terms of t. Nothing can be said about the
character of the motion which is known to be a damped oscillation. The
difficulty encountered is solved when the averaging method is applied. For
brevity of notation we clarify this idea for the example of two first order
equations

(11.5.14)

where jL is a small parameter and functions Ii can be represented as follows

Ii (XI,X2;t) = cp~ (XI,X2)+


L [cp~V) (Xl, X2) cos Avt + 1P~v) (Xl, X2) sin Avt] (i = 1,2). (11.5.15)
v

For instance, the powers and products of cosines and sines in eq. (13) were
replaced by the sums of cosines and sines of frequencies 2AI, 2A2, Al ± A2.
Thus, functions Ii contain components cp~ (Xl, X2) which change slowly in
time (as their arguments Xl, X2 possess this property) and oscillating terms
with slowly varying coefficients cp~v), 1P~v). This prompts one to replace the
right hand sides of eq. (15) by their mean values over a rather large time
interval, Xl and X2 being taken constant during the calculation. Then
T

lim
T-+OCJ
~JIi(XI'X2;t)dt=CP~(XI'X2)
T
(11.5.16)
a
and with this approximation, the equations in (15) are replaced by the
following equations

Xl = jLCP~ (Xl, X2), X2 = jLcpg (Xl, X2). (11.5.17)

For system (13) these differential equations take the form

(8 = 1,2) .

Their solution is given by


604 11. Perturbation theory

and substitution into eq. (8) yields

(11.5.18)

This is still a very rough approximation. It is sufficient to point out that


the term 2h(iIq2 in the expression for the dissipation function which links
the motions is not reflected in solution (18) at all.
A more accurate approximation is obtained in the following way. Return-
ing to system (14) we assume

.= c. ~[ (v) (c C) sinAvt _ oi,(v) (c C) cos Avt] i = 1,2,


x, <",+fJL'I\ <,,1,<,,2 A 'r. <."l,<"2 A
v v v
(11.5.19)

where ~i are solutions of the system of equations (17)

This means that the solution is the sum of the zero-order approximation ~i
and the integral over time of the oscillating terms on the right hand sides
of system (14). During the estimation of this integral Xl, X2 are replaced
by ~1'~2 whose change in time is not taken into account since they change
slowly. By differentiating eq. (19) it is possible to prove that the constructed
solution satisfies the system of differential equations (14) with error of order
fJ2. Details of the construction of higher approximations and substantiation
of the method are given in [12].
Let us carry out the suggested calculation for system (13). By means of
eq. (19) we obtain
11.6 Influence of misbalance on the motion of a heavy top 605

Inserting these expressions into eq. (8) we obtain after simplification

The expression for q2 is obtained under the corresponding replacement of


the indices

The oscillation problem of the dissipative system with two degrees of


freedom about the equilibrium position is studied by Whittaker in [95] by
direct integration of the equations of motion with the dissipative forces be-
ing taken into account. While calculating the natural frequencies Whittaker
neglected the terms proportional to the squares of the coefficients of the
dissipation function and obtained expressions for ql, q2 which differ from
eqs. (21) and (22) only by the underlined terms.

11.6 Influence of misbalance on the motion of a


heavy top
A particle m is attached to a nearly vertical heavy top at distance e from
its axis of symmetry Oz. The perturbed rotation due to the presence of this
606 11. Perturbation theory

unbalance is considered under the assumption that only first order values
of the eccentricity e are used.
The position of the heavy top is described, similar to Subsection 7.18.2,
by the ship angles denoted here as ,\ and J-l (,\ = a and J-l = (3) as well as
the angle of spin <po The coordinates of the attached mass in the coordinate
system Oxyz fixed in the body are denoted by xo = eCOSE,Yo = esinE,zo.
The kinetic energy of the misbalance is

T1 2m
1 [ (WyZO - wzYo) 2 + (wzxo - wxzo) 2 + (WxYo - WyXo) 2]

~m [(w; + w~) z5 - 2wzzo (xow x + YOWy) + ... ] ,


the values proportional to e 2 are not included. Recalling the expressions
for the projections of the angular velocity

Wx -5.cos<p + itcos'\sin<p,
Wy 5. sin <p + it cos'\ cos <p,
Wz tP + itsin'\,
we find that the kinetic energy of the system consisting of the heavy top
and the misbalance is equal to

T = ~ { A (5.2 + it 2 cos2 ,\ ) + C (tP + it sin ,\)2 -


2ezom (tP + it sin'\) [-5. cos (<p + E) + it cos'\ sin (<p + E)]} , (11.6.1)

where A and C denote the equatorial and axial moments of inertia of the
system, respectively.
Designating the unit vector of the upward vertical by k, we construct
the expression for the potential energy

II Q1zck . i3 + mgk· (i1XO + i 2 yo + i3Z0)


(Q1Zc + mgzo) cos'\cosJ-l + mge (k· i1 COSE + k· i2 sinE),

Q1 denoting the weight of the heavy top. Let Q and z~ denote the weight
and the vertical coordinate of the centre of mass of the whole system,
respectively. Since

i1 = ncos<p + n'sin<p, i2 = -nsin<p + n' cos<p,

we have

II = Qz~ cos'\COSfJ - mge [sinJ-lcos (<p + E) + sin'\cosJ-lsin (<p + E)].


(11.6.2)
11.6 Influence of misbalance on the motion of a heavy top 607

Using eq. (1) we obtain

P>. A.\ + mezo (cp + it sin A) cos (tp + c) ,


P/1> P<p sin A + Ait cos 2 A - mezo (cp + it sin A) cos A sin (tp + c) ,
P<p C (cp + it sin A) - mezo [-.\ cos (tp + c) + it cos A sin (tp + c)] .

Therefore, within the adopted accuracy, we have

(11.6.3)

The Hamiltonian function takes the form

H = -12 [p~2 - p<p sinA)2 + J?.


+ (p /1> Acos 2 ACe
1
p2 + 2Q z* cos A cos JL
A
mezo [ P/1> - P<p sin A . ]
AC P>.P<p cos (tp + c) - COSA p<psm(tp+c)-
mge [sin JL cos (tp + c) + sin A cos f-l sin (tp + c)]
(11.6.4)

For a nearly vertical top

mezo .
H = - AC [P>'P<P cos (tp + c) - (P/1> - AP<p) P<p sm (tp + c)]-
mge [f-l cos (tp + c) + A sin ( tp + c)] .
(11.6.5)

The canonical equations of the unperturbed motion of the top

. P>.
A=A'

it = ~ (P/1> - P<pA) ,
608 11. Perturbation theory

reduce to the form

(11.6.6)

Denoting the roots of the quadratic equation

(11.6.7)

by qI, q2 and the initial values of the coordinates and momenta by a", a>., {JI'
and {J >. we come to the equalities

(11.6.8)

It follows from the last approximated equality in (6) that

(11.6.9)

where a<p denotes the initial value of 'P. Without loss of generality we can
assume c = 0 and then the perturbing function is as follows

HI = ;~~ i{J<p [(PI' - P<p).. + ip>.) e-i<p - (PI' - P<p).. - ip>.) ei<P] -

~mge [(p, + i)..) e-i<p + (p, - i)..) ei<P] = HI (aI" a>., a<p, {JI" {J>., (J<p; t) .
(11.6.10)

The variables aI" ... ,{J<p imply the initial values of the coordinates and
momenta. For this reason, the equations for the perturbed motion have
the canonical form (2.3). These equations can be easily constructed by
replacing the coordinates and momenta in eq. (10) due to formulae (6) and
(8). Further analysis will be limited to the case of a rapidly rotating top,
that is when the condition
11.6 Influence of misbalance on the motion of a heavy top 609

holds true and the roots of quadratic equation (41) are approximately equal
to

(11.6.11)

The first root corresponds to a slow motion known as precession whereas


the second root describes fast nutational jitter of the axis of the top.
Under the above conditions formulae (8) can be simplified by taking

fL + iA = + io:.>-) eiq,t + A~2 ({3JL- (3<p0:.>- + i{3.>-)


(O:JL (e iq,t - e iq2t ) , }
PJL - P<pA + ip.>- = iAqI (O:JL + io:.>-) (e tq,t - e tq2t ) +
({3JL - (3<p0:.>- + i{3.>-) e iq2t .
(11.6.12)

Then HI can be represented as follows

HI = mezo [
2AC i{3<p iql (O:JL + io:.>- ) (etq,
. t .
- e tq2
t) e-t<p+
.

iqI (O:JL - io:.>-) (e- iq,t - e- iq2t ) ei<p +~ ({3JL - (3<p0:.>- + i{3.>-) ei (q2 t -'P) -

~ ({3JL - (3<p0:.>- - i{3.>-) e- i (Q2 t -<P)] - ~mge [(O:JL + io:.>-) e i (q , t-<p)+


(O:JL - io:.>-) e- i (q, t -<p) + A~2 ({3JL - (3<p0:.>- + i{3.>-) (e iq,t - e iq2t ) e-i'P_

A~2 ({3JL - (3'P0:,>- - i{3.>-) (e- iq,t - e- iq2t ) ei<p]. (11.6.13)

Proceeding to construction of the canonical equations of perturbed mo-


tion

and their complex conjugates (1' (2. Hence,


610 11. Perturbation theory

which enables the above equations to be cast in the form

(11.6.14)

The remaining two equations are

i3cp -- - 8H
8acp -_ - 8H
1 1
8<p , }

(11.6.15)
. _ 8H1 _ (8 H l) _ QZ~ 8H1 .!. 8H1 -!i3
acp - 8f3cp - 8f3cp f3! 8ql +A 8q2 C cpo

The first of these determines the perturbation of the kinetic moment. While
constructing the second equation it has been taken into account that Hl
depends on f3cp both explicitly and through ql, q2, (p. By virtue of eqs. (9)
and (15) we find that the change in the angular velocity of spin due to the
misbalance is

~. . f3cp . t 13' (8H1 ) Qz~ 8H1 18H1 ( )


<p = <p - C = acp +C cp = 8f3cp - f3! 8ql +A 8q2' 11.6.16

As follows from eq. (13) this change is a superposition of oscillations with


frequencies

Returning to equations (14) we can simplify them by accounting for the


smallness of ql compared with q2. Quantities f3cp and acp on the right hand
sides of these equations are considered as constant values with f3cp/C = (Po.
Introducing the small parameter x = rr;~o we obtain

(11.6.17)
11. 7 Rotation of an Earth satellite about its centre of inertia 611

It follows from these relationships that if A = C, i.e. the inertia ellipsoid is


a sphere, the expressions for the perturbation contain terms which increase
linearly with time.

11.7 Rotation of an Earth satellite about its centre


of inertia
As shown in Sec. 5.5 the forces of the central field acting on the moving
body are reduced not only to the resultant force F but also to the resultant
moment rn G about the centre of inertia of the body. Under the assumption
that the central ellipsoid of inertia of the body is an ellipsoid of revolution,
the expression for the moment rn G is given by formula (5.5.13). When mo-
ment rn G is absent the rotation of the body about its centre of inertia is
a regular precession. The regular precession implies that the body rotates
with a constant angular velocity about the rotation axis of the ellipsoid
of inertia and this axis in turn rotates with a constant angular velocity
about a fixed direction of the vector of the principal angular momentum
KG, the angle between them being constant. The problem of perturbation
of the regular precession caused by the presence of moment rn G belongs
to a number of the classical questions of the celestial mechanics explaining
the phenomena of lunisolar precession of the earth axis and equinox anti-
cipation. Rotation of the satellite which is a body extended along its axis
of symmetry about its centre of inertia presents a perturbed regular pre-
cession too. Consideration of this problem provides us with an interesting
example of the application of perturbation theory. Elementary analysis of
the lunisolar precession is given in [31]. This problem, as well as the prob-
lem of lunar libration, is set out at great length in the classical treatise on
celestial mechanics [89]. The satellite precession is outlined in [4] where the
perturbed action of the gravitational and aerodynamical forces, as well as
variation of the orbit of the centre of inertia of the satellite are studied.
Let us denote the axial and equatorial moments of inertia of the satellite
by C and A, respectively. Its motion occurs in the field of the attraction
force of the earth which is considered here as a homogeneous sphere. Then
the potential energy of the attraction force, due to eq. (5.5.6), is equal to

II = -J-l [~o + 2~5 (C - A) (x 2 + y2 - 2z2)] ,

where J-l = f M, f and M denoting the universal gravitational constant


and the mass of the earth, respectively. Replacing x 2 + y2 by r2 - z2 and
cancelling out in II the terms which are independent of the orientation of
the axes bound to the satellite, we obtain
3 z2
II = 2J-l (C - A) r 5 · (11.7.1)
612 11. Perturbation theory

Let e r and ia denote the unit vector of the direction MG from the centre
of the earth to the centre of inertia of the satellite and the unit vector of
the axis Gz of the inertia ellipsoid, respectively, see Fig. 5.4b. As the angle
between these vectors is (), we have

Z2 • 2 2
2" = (er ·13) = cos (),
r
and the expression for II takes the form

(11. 7.2)

Let us introduce into consideration the orthogonal trihedron of the unit


base vectors el, e2, e3 with origin at the centre of the earth which is a focal
point of the elliptic orbit of the satellite. Vector el is directed to the orbital
perigee, e2 lies in the orbital plane and is directed to the side of motion
from the perigee to the apogee and e3 = el x e2 is perpendicular to the
orbital plane. The perturbation of the orbital elements is neglected, then
vectors el, e2, e3 remain fixed in the space. Notice also that

(11. 7.3)

where a denotes the true anomaly that is the angle between vector e r and
the direction to the perigee.
The orientation the unit vectors it, i 2 , ia of the axes Oxyz bound to the
satellite is given by the Euler's angles 'ljJ, {), cp. Then

el . i3 = sin {) sin 'ljJ, e2· i3 = - sin {) cos 'ljJ, e3 . i3 = cos {) (11.7.4)

and moreover, due to eq. (3)

cos {) = e r . ia = sin {) (sin 'ljJ cos a - cos 'ljJ sin a) = sin {) sin ('ljJ - a) .
(11. 7.5)

Denoting now the major semi-axis of the orbit by a and recalling eq.
(10.1.5.15) we obtain

3 a3
II = '2n2 r3 (C - A) sin2 {)sin2 ('ljJ - a), (11. 7.6)

where n denotes the mean motion of the satellite.


The expression for the kinetic energy of the rotational motion of the rigid
body about its centre of inertia has the form
11.7 Rotation of an Earth satellite about its centre of inertia 613

It can be cast as follows

then, noticing that C - A < 0 for the body of revolution extended along
axis Gz, we obtain

C
T
IIII
=
3n 2 a3
-23" (
1- A
w r 1_ 1_ _
C) ')'2
3n 2 a3 (A
sin2'!9sin2(~-a) < -23"
W r
C )
-1 .
A
(11.7.8)

The ratio (nlw)2 is small for satellites which enables the application of
perturbation theory, the first approximation (the auxiliary problem) cor-
responding to the regular precession in the case II = O. By virtue of eq.
(5.6.34), we have for the earth

-----c-
C-A
= 0.0033.

For the solar attraction


n 1
w 366

and for the lunar attraction


n 1
:; ~ 28

Thus, the ratio IIII IT proves to be very small. For the satellite the rela-
tionships turn out to be less favourable from the perspective of applying
perturbation theory.
It follows from expressions (6) and (7) that <p is a cyclic coordinate, thus
the corresponding generalised momentum

(11. 7.9)

retains constant value in both initial and auxiliary problems.


By constructing the expression for the momenta

(11.7.10)

we find
. Pi}
'!9=A' (11.7.11)
614 11. Perturbation theory

The Hamiltonian function takes the form

H
1 p2
= - [
....:!!.. + (P'1f; - f3 <.:Os {}
<p )2] + _n
3 a3
2_ (C - A) sin 2 19sin2 ('I/J - (J)
2 A A sin 2 19 2 r3 '
(11. 7.12)

where the constant term f3~/2C is omitted. The Hamiltonian function for
the auxiliary problem equals

Ii = ~ [p~ (P'1f; - f3<pCOS19)2] (11.7.13)


o 2 A + A sin 2 19 '

which corresponds to the regular precession, as pointed out above. The


coordinate 'I/J is cyclic, hence

(11.7.14)

The vector of the angular momentum of the body can be written in the
form

K = A ( m9 + n' ~ sin 19) + Ci3 (<p + ~ cos 19) ,

where n, n', h denote, as usual, the "half-moving" trihedron of the unit


vectors of the nodal line, the perpendicular to it in the equatorial plane
and axis Gz, respectively. Furthermore
w n ={}, wnl=~sin19, w3=<p+~cos19
denote the projections of the angular velocity vector on these directions.
By virtue of eqs. (9), (11) and (14) we obtain
,P'1f; - f3<p cos 19 .
K = np'!') +n
. 19
sm
+ 13f3<p. (11. 7.15)

Since the moment of the external forces about the centre of inertia is
zero, i.e. m G = 0, the vector K remains constant in both its value and
direction. This follows from the theorem on change in angular momentum.
However one can prove the above said without referring to this theorem.
To this aim, we differentiate expression (15) for vector K and use only the
canonical equations of motion (11), (14) and the remaining equation

(11. 7.16)

Let us recall that the vector of angular velocity of the" half-moving" trihe-
dron of unit vectors n, n', i3 is given by
Wo n{} + n' ~ sin 19 + i3~ cos 19 (11. 7.17)
1 ( ,f3'1f; - f3<pcos19 • f3'1f; - f3<pcos19 )
A np'!') + n . 19
sm
+ 13 sm . 2
19
cos 19
11.7 Rotation of an Earth satellite about its centre of inertia 615

and thus
• _ _ f3,p - f3cpcos<O ( ,cos<O _. )
n - Wo x n - A . 'o n..o 13 ,
SIn 'If SIn 'If
f3,p - f3 cos <0
piJ
h- cos <0, (11.7.18)
il' = Wo x n' = A - n
. CP2
Asm <0
di3 • ,PiJ f3,p - f3<p cos <0
di=W o xI 3 =-n"A+ n A sin <0 .
We find
• _. • (f3,p-f3cpcos<o)-, f3,p-f3cpcos<O., dh_
K -'piJn + piJn + . <0 n + . <0 n + f3cp dt - 0,
sm sm
(11.7.19)

which can be easily proved by inserting expressions (11) and (16) for the
derivatives -0, piJ and the unit vectors. As follows from eq. (15)

(11. 7.20)

The first equation indicates that the body axis Gz retains a constant angle
with a fixed direction of vector K. With this in view, if we direct, for
instance, vector e3 along K and denote the corresponding Euler angles by
w,8,<I> then we obtain, due to eq. (20), that

8 = 8 0, f3w = K, f3q, = K cos 8 0, (11.7.21)

and, by meanS of eqs. (11) and (9), that

"if! = f3 w - f3q, cos 8 0 = K . f3q,. A - C


<I>=C-wcos80= AC Kcos8 0 ·
Asin 2 8 0 A'
(11. 7.22)

These expression give the angular velocity of axis Gz in the case of regular
precession about the fixed direction of vector K and the angular velocity
of the spin, respectively.
Let k denote the unit vector of the direction of vector K. According to
definition (3.1.12) the vector of finite rotation () of axis i3 at time instant
t is
W Kt
() = 2ktan"2 = 2ktan 2A' (11. 7.23)

provided that the initial position is ig at t = O. By virtue of Rodrigues's


formula (3.1.11) we obtain

h = i~ + 1 W [2 tan Wk x i~ + 2tan2 Wk x (k x i~)]. (11.7.24)


1+~~- 2 2
2
616 11. Perturbation theory

Taking into account that K is a fixed vector we obtain, by means of eq.


(15),
_ ~ [ 0
k - K n (3{) + n ,0 (3'IjJ - (3'1' cos {)o
. {)
sm 0
+ 13(3<p
.0 ]
, (11.7.25)

where (3<p denotes the initial value (at t = 0) of momentum p{) and the
index zero denotes the initial values of the other quantities. Carrying out
calculations we have
i3 = ig cos \If + k x i~ sin \If + (1 - cos \If) kk . i~
or

i3 n
o [(3'IjJ - (3<p cos{)o sin \If
. {)
(3<p(3{) (1
- K + -K2 - cos 'l'
'T')] +
sm 0

n'o [_(3/i~\If + (3<p ((3~;s~:{):os {)o) (1 - cos \If)] +

ig [cos \If + ~ (1 - cos \If)]. (11. 7.26)

By means of this relationship it is easy to obtain the expressions for Euler's


angles {) and 'l/J in the case of regular precession as functions of time and
their initial values. To this end, it is sufficient to project vector i3 on the
axes of the fixed direction. Recalling the formulae
h . el = sin {) sin 'l/J, i3 . e2 =- sin {) cos 'l/J , h . e3 = cos{),
,0 . ./. {)
nO. el = cos'l/Jo, n . el = - sm '1-'0 cos 0; nO. e2 = sin'l/Jo,
,0 = 0, ,0 . {)
n . e2 = cos'l/Jo cos{), nO. e3 n . e3 = SIn 0,
we arrive, after elementary transformations, to the following expressions
(3 (3'IjJ) (3{) sin {)o . (3'IjJ(3
cos {) = cos \If ( cos {)o - ~2 - K sm \If + K2 <p , (11. 7.27)

sin \If ((3'IjJ - (3 cos {)o


sin {) sin 'l/J = sin {)o sin 'l/J o cos \If + - - . <p{) cos'l/J o+
K sm 0
1 - cos \If ( (3<p - (3'IjJ cos{)o )
(3{)cos{)osin'l/Jo)+ K2 (3<p (3{)cos'l/Jo+ sin{)o sin'l/Jo,
(11.7.28)

(11.7.29)
11. 7 Rotation of an Earth satellite about its centre of inertia 617

The first of these formulae is obtained by integration of the energy integral,


whilst the second and the third by integration of expression (11) for the
angular velocity;P. Of course, this derivation, based on the simple formulae
for the regular precession, is essentially more simple. Let us recall also that

(11.7.30)

In addition to formulae (27)-(29) we write down the relationships

P.p = f3.p, P19 = Si~1} [(cos1}o - f3~.p ) sin \(T + i; sin 1}0 cos \(T] ,
(11.7.31)

the second one being obtained by differentiating expression (27).


Thus, we obtained the generalised coordinates and momenta for the reg-
ular precession which are expressed in terms of time and initial values of
the variables. This means that the solution of Cauchy's problem for the
auxiliary system of canonical equations is obtained. For this reason, the
equations of perturbed motion for quantities f319, f3.p, 1}0, 'l/Jo are canonical.
Due to eqs. (11.2.3) and (12) these equations are

. all . all . all (11.7.32)


1}0 = af3 19 ' 'l/Jo = af3.p' f3.p = - a'I/J·

We will not analyse these equations and turn our attention to another
means of calculation. To begin with, we consider the resultant angular
moment of the initial problem rather than the auxiliary problem. Replacing
f3.p in eq. (15) by P.p (this momentum is no longer constant) and using
formula (12) we have
3
K2 = 2AH + f3~ - 3n2 a3 A (C - A) sin 2 1} sin 2 ('I/J - a) . (11. 7.33)
r

In as much as the energy integral exists we can write

. • 2 d a3 2 2
2KK = 2K· K = -3n A (C - A) dt r3 sin 1}sin ('I/J - a). (11.7.34)

The canonical equations of the initial problem can be written in the form
618 11. Perturbation theory

where (p-a)o is given by eq. (16). Equations (11), as well for formulae of
differentiation (18) do not change when {3,/} is replaced by p",. Hence, taking
the time-derivative of vector K we obtain instead of (19)

3
K = -3n2a3 (C - A) sin -0 sin (¢ - a) [ncos-osin (¢ - a) + n' cos (¢ - a)] ,
T
(11. 7.36)

since the result of the other differentiations is identically equal to zero due
to eq. (19). We obtain

3
K . el = -3n2 a3 (C - A) cos -0 sin -0 sin (¢ - a) sin a,
T3
K·e2 = -3n2a3 (C-A)cos-osin-osin(¢-a)cosa, (11.7.37)
T3
K· e3 = -3n 2a3 (C - A) sin 2 -0 sin (¢ - a) cos (¢ - a).
T

In what follows, our analysis is limited to consideration of only secular


perturbation of the regular precession, i.e. the terms whose values increase
monotonically with time. All periodic perturbations are ignored. Let us
notice that the right hand sides of eq. (37) contain two groups of periodicity,
namely periodicity with respect to argument a with the period of revolution
of the satellite on the orbit and that with respect to argument l}i with
the period of precession of the satellite axis, the latter being smaller than
the first. It is necessary to average the right hand sides of the mentioned
equations over the arguments a and l}i taking into account neither long-
periodic nor short-periodic perturbations. Then we come to relationships
of the following kind

J J
211" 211"

K· el = ~3n2
47f
(C - A) a: sin ada
T
sin -ocos-osin (¢ - a) dl}i.
o 0
(11.7.38)

In order to gain a more precise picture of the perturbed motion it is nec-


essary to take into account the long-periodic oscillations restricting the
analysis by averaging only over variable l}i and neglecting the change in
the quantities depending on a since they do not change considerably over
the period of variable l}i.
As follows from relationship (34) the value of vector K is constant even
after this averaging. Averaging the right hand side of eq. (34) over l}i we
11. 7 Rotation of an Earth satellite about its centre of inertia 619

obtain

2KK .
= -3n 2 A (C - A) 27f 1 J
271"
d a3
d\J! dt r3 sin 2 'l9sin 2 ('¢ - a)
o
3 2 K 3/ W=271"
= -~A (C - A) - sin2 'l9sin2 ('¢ - a) a3 = 0, (11.7.39)
27f A r w=o
since sin '¢ sin 'l9 and cos'¢ sin 'l9 are periodic functions of \J! due to eqs. (28)
and (29).
Let us proceed to estimate the integrals of type (38). While averaging it
over a it is necessary to use eq. (10.15.20) and replace a 3 j r 3 by

1
(1 + e cos a) .
3
------".3 (11.7.40)
(1 - e 2 )
It does not complicate the integration, however the notation becomes bulky.
With this in view, the forthcoming calculation is carried out for a circular
orbit (e = 0). This does not lead to a considerable quantitative error as the
existing satellites have small eccentricity. The averaging over a yields

. 3n2 J271"
K· el = - - (C - A) cos'l9sin'l9cos'¢d\J!,
47f
o (11.7.41)
J
271"

K . e2 = - 34: (C - A) cos 'l9 sin 'l9 sin '¢d\J!


o
and

J J
271" 271"

K . e3 = - ::: (C - A) sin 2 'l9d\J! sin 2 ('¢ - a) da = O. (11. 7.42)


o 0

The projection of vector K on the normal e3 to the plane of the cir-


cular orbit proves to have no secular perturbations. The integrands in eq.
(41) should be replaced by the corresponding expressions (27)-(29). While
estimating the integrals the quantities (31/)) (3{h 'l9 0 , '¢o should be taken as
constant values as they are slowly varying functions of time. Then we have
620 11. Perturbation theory

where the terms yielding zero after integration over \[I are omitted.
Having carried out the integration and further simplification, we have

. 3n
K·el=-(C-A) (3tJ~)
2
1-- -
tJ,;; (tJ<p - tJ,;;. cos{)o cos7/Jo--sm7/Jo
tJiJ.)
4 K2 K Ksm{)o K
(11.7.43)

and by analogy

. 3n 2
K·e2=-(C-A) 1 - -
(3tJ~) -tJ,;; (tJ<p - tJ,;;. cos {)o sm7/Jo+-cos7/Jo
. tJiJ )
4 K2 K Ksm{)o K

These relationships can be easily cast in a more transparent form. By using


eq. (25) we find

(11.7.44)

In the case of the regular precession we take into account eq. (20) and
denote the constant angles of vector K with the satellite axis and the
perpendicular to the orbital plane by eo and Xo, respectively. Then we
obtain

.
K· el = - - -
2
3n
CC- A (
cose o 1- 3 cos 2)
eo cosXoK· e2, }
4 ro
(11.7.45)
.
K· e2 = 4
2
3n
C- A ( 2)
Cro coseo 1- 3 cos eo cosXoK· el,

where ro is the constant angular velocity of spin given by eq. (9). These
relationships show that the vector of the angular momentum K rotates
with angular velocity

n=
3n A
2 C - 2
- - C cos eo (1- 3 cos eo) cosXoe3, (11.7.46)
4 ro

directed parallel to the normal e3 to the trajectory plane, [4]. Thus, the
satellite executes a regular precession about constant vector K which pre-
cesses with angular velocity n. The effect of perturbation on the regular
precession is explained by the non-sphericity of the inertia ellipsoid of the
satellite (A -=I- C).
11.8 Equations of the perturbed Keplerian motion 621

11.8 Equations of the perturbed Keplerian motion


The Keplerian motion, which is the motion of a particle in the field of the
central force of attraction

(11.8.1)

is considered in Sec. 10.15. One of the basic problem of celestial mechanics


is calculation of the perturbations of the Keplerian motion caused by action
of accessory forces which are small compared with Fa. The problem of per-
turbation of motion of the moon due to solar gravitation, being a particular
case of the celebrated three-body problem, is one of these problems. Also
motion of the satellite is perturbed by resisting force of the atmosphere
and additional gravitational forces due to oblateness of the earth.
Let us turn our attention to derivation of the perturbed Keplerian mo-
tion. A derivation of the equation of the perturbed Keplerian motion based
on an elegant calculation of the Lagrange brackets for the elliptic elements
of the orbit was suggested in [31]. Geometric constructions which are diffi-
cult to follow are given in [52] and [24]. The present section is based on the
author's paper [59]. We start with the expressions (10.15.21) and (10.15.24)
for the radius vector r and the velocity vector v of a planet in the unper-
turbed motion. Expressing the eccentric anomaly w in these equations in
terms of the true anomaly 'P by means of eq. (10.15.18) we obtain the
following expressions

v = I¥
- ~
a 1-
1
e2
.
[ere sm 'P + e<p (1 + ecos'P)].
(11.8.2)

Here e<p denotes the unit vector of the perpendicular to the radius vector r
directed to the side of increasing 'P and lying in the plane of the unperturbed
motion, and e3 = e r x e<p' The trihedron en e<p, e3 is obtained by rotation
of the trihedron el, e2, e3 about e3 through angle 'P, with el being the
direction from the attracting centre to the pericentron of the unperturbed
orbit. As mentioned in Sec. 10.15 the orientation of latter trihedron with
respect to the fixed axes is given by Euler's angles n, i, w which are the
longitude of the ascending node, the orbit inclination and the angle of the
line of nodes with the direction to the pericentron.
Following the idea of parameter variation we use the same notation (2)
for the vectors r and v in the perturbed motion as in the unperturbed
motion, however the elliptic elements (10.15.22) are taken to be sought-for
functions of time. They should be determined in such a way that differ-
entiation of r with respect to time would lead to the above expression for
v and differentiation of v would yield the acceleration vector given by the
geometric sum of the attraction force Fa and perturbing force F.
622 11. Perturbation theory

Let w denote the vector of the angular velocity of the trihedron el, e2, e3,
then the angular velocity of the trihedron e r , ecp, e3 is w + e3CP. According
to the formulae for differentiation of unit vectors (2.7.5) we obtain

=:
~r (w + e3~) x e r ~ (W3 + cp) ecp - ~cpe3' }
ecp - (w + e3CP) x ecp - wre3 - (W3 + cp) en (11.8.3)
e3 = (w + e3CP) X e3 = wcpe r - wrecp.

Here wr , wcp, W3 are given by formulae (2.9.3) expressing projections of w


in terms of the derivatives of Euler's angles

. cP = di
Wr = WI cos cP + W2 sm dt co~ u + ;,.
~ ~ sm z. sm
. u, }
. dz.;, . . (11.8.4)
Wcp = -wIsmcp+w2coSCP = --d smu+~~slnzcosu,
. t
W3 = W + Ocosi,

where

(11.8.5)

denotes the angle of the radius-vector of w with the direction on the as-
cending node.
In the unperturbed motion the trihedron e r , ecp, e3 has angular velocity
e3cpo which is different from the component e3CP of the vector of angular ve-
locity of this trihedron in the perturbed motion. Hence, in the unperturbed
motion formulae (3) become

(11.8.6)

Differentiating expression (2) for vector r with respect to time yields

f =

(11.8.7)

the two first terms implying the derivative fo calculated for the unperturbed
motion, and they give value v defined by formula (2). This is easy to prove
if we replace cpo by

(11.8.8)
11.8 Equations of the perturbed Keplerian motion 623

determined byeqs. (10.15.3) and (10.1.5.14) (this can be obtained directly


from the theorem on areas). Due to above, the additional terms in formula
(7), as well as the coefficients of e3, ecp, e r should vanish. This yields the
first set of equations for the perturbed motion

di.
wcp=-dtsmu+~,smzcosu=
r... 0
, (11.8.9)

cp = cpo _ W3 = cpo - (w + ncos i) , (11.8.10)

w3esincp + ~2
1-e
(2e + cos cp + e2 coscp) - ~a (1 + ecos cp) = 0, (11.8.11)

where in eq. (11) value cp - cpo is taken from eq. (10).


Let us proceed to differentiation of vector v. Using eqs. (6) and (8) we
construct first the expression for va which is the derivative of v in the
unperturbed motion

va av.o
= -cp
acp
+ e3CP.0 x v

(11.8.12)

Next, we construct the equality

v - vO = F = Fre r + Fcpecp + F3e3. (11.8.13)

Because of eqs. (9) and (10), the expression for w + e3CP takes the form

(11.8.14)

eq. (13) is cast as follows

. . 0 avo avo avo ( .0) av. .


v - v = -cp + - e + - a + wre r + cp e3 x v - -cpo - cp oe3 x v


acp ae aa acp
. a ) v+w r e3
a . a +a-a 1 (1 +ecoscp,
)
a cp- +e-
= ( -W3 ae a
- ~
a v 1- e 2
(11.8.15)

and the second set of equations for perturbed motion becomes

Wr
di cos u +
= -d ~,
. u= ~
. Z. sIn
r. SIn - JI=e2 F 3, (11.8.16)
t 1-l1+ecoscp
624 11. Perturbation theory

e(cos'P+e)
----'----'--::-2--'- - -
. =
a (1 + ecos'P) + W3esm'P ~ ~2
- y 1- e F<p, (11.8.17)
1- e 2a fl

_e_'_ sin'P - ~esin'P - W3 e cos 'P = ~~Fr. (11.8.18)


1 - e2 2a Vfl
Equations (9) and (16) yield

-di = ~~
- F 3 cosu, o
. sin i = ~~
- F3 sin u,
dt fl 1 + e cos 'P fl1 + e cos'P
(11.8.19)

and eqs. (11), (17) and (18) render

e. -_ ~- y r;--;)1
2
1 - e~
(F'sm 'P + e +
r
2 1cos 'P + e cos 2 'P F<p ) , (11.8.20)
fl + ecos'P

-a = ~ 1
- ~ .
[Fresm'P + (1 + ecos'P) F<p], (11.8.21)
2a fl2 1- e

W3 = ~-fl v'I-e2
e
(
- Fr cos 'P + 2+ecos'P. )
F<p sm 'P .
1 + ecos'P
(11.8.22)

The sixth equation of perturbed motion for the time of pericentron passage
to will be obtained in Sec. 11.9. The derivation is based on relationship (10).
Taking into account the value of 0 given by the second equation (19) and
formulae (4), (5) and (8), relationship (10) is transformed to the form

. = ~(1 + ecos'P)2 r
u V~ 3'
(11.8.23)
a a (1 - e 2 )"2

Using this expression we transform to the independent variable u in eqs.


(19)-(22) and obtain the equations for perturbed motion in the form

di = G F3 COSU dO = G F 3 sinu
(11.8.24)
du 1 + e cos 'P ' du (1 + e cos 'P) sin i '

de _ G
-d -
(F'rsm'P+ e + 2 1cos 'P + e cos 'P F<p ) ,
2
(11.8.25)
u + ecos'P

1 da G
--d = --2 [Fr e sin 'P + (1 + e cos 'P) F<p], (11.8.26)
2a u 1- e
11.9 Perturbed motion of the centre of inertia of the Earth satellite 625

dw = -G ( - F r cos cp + 2 + e cos cP F .sm cp - eF 3 sin u cot i )


- . (11.8.27)
du e 1 + e cos cp <p 1 + e cos cp
Here for the sake of brevity
1a2 (1_e 2 )2
G = - - ----'--.!...--,;- (11.8.28)
r JL (1 + e cos cp)2 ,
and eq. (27) is obtained from (22), (24) and the last equation in (4). The
present derivation of the equations for perturbed motion not being re-
lated to the geometric constructions requires uncomplicated calculations
suggested by the method of parameter variation.
Angle cp in eqs. (23)-(26) is determined due to eq. (5), as the difference
of angles u and w. If F r , F<p, F3 do not depend explicitly on time, then the
five equations (24)-(28) from the system of six equations (23)-(28) become
independent of eq. (23). After eqs. (24)-(28) have been integrated, eq. (23)
yields an expression for time in terms of the variable u.

11.9 Perturbed motion of the centre of inertia of


the Earth satellite
The present analysis is limited to consideration of the perturbation of the
Keplerian motion of the centre of inertia of a satellite caused by a non-
central field of the earth attraction. Paper [73] is devoted to the influence
of the atmosphere resistance. The secular perturbations due to the fact that
the gravitational field is not central are studied there also. The simultaneous
influence of these factors is analysed in [87]. The problem of the satellite
motion in a non-central force field is solved in [8] and [9] by means of
approaches different from the classical perturbation technique.
Substituting 8 3 = C and 8 1 = 8 2 = A into eq. (5.5.6) yields the
potential energy of the additional force of attraction due to the oblateness
of the geoid

The origin of the coordinate system D~7]( is taken at the centre of the earth,
axis D( being perpendicular to the equatorial plane. Replacing C - A in
eq. (5.6.23) we obtain

(11.9.1)

where c = 0,00164 is a constant of the earth shape. Its smallness leads to


a slow variation of the orbital elements. The values proportional to c 2 in
eq. (1) are not considered here and in the sequel.
626 11. Perturbation theory

We have

P,H5 ( 1- 5-
F = -gradII = -E:-- (2 ) e
r -
P,H5 (
2E:---k, (11.9.2)
r4 r2 r4 r

with k denoting the unit vector of axis 0(. Noticing that

-(=k . e r = SIn'/,
. . SIn
. u, k . e", = sin i cos u, k· e3 = cosi, (11.9.3)
r
we find

Fr = _E:P,~6
r
(I-3sin 2 isin 2 u),

P,H5.2"2 (11.9.4)
4- SIn '/,SIn u,
= -E:-
D
r",
r
p,R6
F 3 = -E:- . 2' .
4- sm '/, SIn u.
r
The value r is given by

r = 1 + 2E:
H5 cos2 i 2 (1 + ecoscp) sin2 u. (11.9.5)
a2 (1 - e2 )
It will be needed later, however we can take r = 1 for construction of the
equations for perturbed motion (8.24)-(8.27). Inserting expression (4) for
forces into these equations we arrive at the following system of equations

di = -E:I.
-d - sm 2''/, (1 + ecoscp ) SIn
. 2u, (11.9.6)
u 2

dO. .
du =-E:Icos'/,(I+ecoscp)(I-cos2u), (11.9.7)

de
-E:l [(1- 3sin2 usin 2 i) sincp (1 + e cos cp)2 + (11.9.8)
du
sin 2 i sin 2u (e + 2 cos cp + ecos 2 cp) (1 + e cos cp)] ,

1 da
2adu
(11.9.9)

dw
E:: [(1- 3sin2 usin 2 i) (1 + e cos cp)2 coscp-
du
sin 2 i sin 2u sin cp (2 + e cos cp) (1 + e cos cp) +
2e sin 2 u cos 2 i (1 + e cos cp)] , (11.9.10)
11.9 Perturbed motion of the centre of inertia of the Earth satellite 627

where

(11.9.11)

Restricting our consideration to determination of the secular perturba-


tion only we replace the right hand sides of these equations by their mean
values over one period of revolution of the satellite, i.e. over the interval
(0,21f) of the change of variable u. While averaging we deem the elements
a, e, i, w of the elliptic orbit to be constant whereas angle cp should be re-
placed by u - w. For example, let us consider the process of averaging the
value in the square brackets in eq. (8). We rewrite it in the form

(1- ~ sin2 i + ~ cos 2u sin 2 i) [sin cp + e sin 2cp + ~e2 (sin cp + sin 3CP)] +
sin 2 i sin 2u [~e + (~e2 + 2) cos cp + ~e cos 2cp + : cos 3CP]

= ~e sin 2 i (cos 2u sin 2cp + sin 2u cos 2cp) + ...

= ~esin2 isin (4u - 2w) + ...


The dots denote terms whose mean value is zero. A similar calculation
yields

da = 0 de = 0 di
du = O. (11.9.12)
du ' du '
Therefore, the elliptic elements a, e, i are periodic functions of u containing
no secular (i.e. proportional to u) terms. The mean values of the derivatives
dO . dw 1 ( 1-5cos 2 Z.)
- = -E1 cosz -=--E1 (11.9.13)
du ' du 2
turn out to be non-zero. During N revolutions of the satellite the ascending
node and the angle between the direction to the pericentron and the line of
nodes gain respectively the following values

(11.9.14)

The change in the sixth elliptic element, which is the time of pericen-
tron passage, is much more elaborate. Let us notice first that relationship
(10.15.17) for the true anomaly cp, in terms of the eccentric anomaly w,
is an integral of the equations for the unperturbed motion containing the
three constants e, a, to (the latter two constants are due to Kepler's equa-
tion (10.15.16)). Hence, in accordance with the basic idea of the method
of parameter variation, the form of integral (10.15.17) retains a constant
value throughout the perturbed motion for non-constant e, a, to.
628 11. Perturbation theory

Bearing this in mind and differentiating eqs. (10.15.17) and (10.15.17)


with respect to time we obtain

. ~2 tV esincp . (+ esinw
cp - V 1 - e-
- 1 - e cos w
+- -
1 - e2 '
w= , (11.9.15)
1- ecosw
where ( = n (t - to).
Removing tV from these equations and replacing cos wand sin w by eq.
(10.15.18), we obtain

. _ ((1 + e cos cp)2 e sin cp ( )


cp - 3/2 + 2 2 + e cos cp . (11.9.16)
(1 - e 2 ) 1- e
By virtue of eq. (8.23)

if! = U-W=U(l- dwdu )= Vfi!.(1+ecoscp)2r(1_


-;;: a (1 - e2)3/2 du
dw )

~ fi!.(1 + ecoscp)2 [1 + (r -1) _ dW] (11.9.17)


V-;;: a (1 - e 2)3/2 du
and substitution into eq. (16) yields

( _ n =n [r _ 1 _ dw _ de sin cp (2 + e cos cp)] . (11.9.18)


du du 1- e 2
By means of eqs. (5), (10) and (8) the right hand side of this equation is
reduced to the form

nCl (1 e~le ~o:2~)3 [(1 - 3 sin 2 u sin 2 i) (2e - cos cp - e cos 2 cp) +
sin 2 i sin 2u sin cp (2 + e cos cp) ] . (11.9.19)
On using relationship (23) and Kepler's equation, the expression on the left
hand side is transformed as follows

(-n = n. (t - to ) - ndto
- (11.9.20)
dt
[.!.n (w _ esinw) dn
du
_ n dt o ] n (1 + ecoscp)2
du a (1 _ e 2)3/2
Noticing that
1 dn 3 da
(11.9.21)
n du 2a du
and replacing wand sin w by means of eqs. (10.15.17) and (10.15.18) we
come to the following differential equation

nddt- = II (cp) 2 arctan ~-e cp + h


- - tan -2 (cp), (11.9.22)
u l+e
11.9 Perturbed motion of the centre of inertia of the Earth satellite 629

in which

h ('P) -3CI
-2
1- e
(1 + e Cos 'P )2 [(1 - -3.
2

sm z + -3 COS 2u sm
2
2') .
. z e sm cp+

(1 + ecos 'P) sin 2 isin2ul ' (11.9.23)

12 (cp) ~ (1 + ecos'P) {(1- ~2 sin 2 i + ~2 cos2usin 2 i)


e 1 - e2
X

[e (2 + e 2 ) - cos'P (1- e 2 ) - e (1 + 2e 2 ) cos 2 'Pl +


sin 2 i sin 2u sin 'P [2 + e 2 + e cos 'P (1 + 2e 2 ) 1}. (11.9.24)

The mean value of function 12 ('P) over the period of the satellite revolution
is easy to calculate

-1 J271"
12 ('P) d'P = - ~
CI [( 1 - -3. e 2 ) +""8
sm 2 z.) ( 1 + 2 3e 2 sin 2 i cos 2w ] .
27f vI - e 2 2
o
(11.9.25)

Function h ('P) is represented by the trigonometric series

5
h ('P) = 3CI 2 ~ (ak sinkcp + bk cosk'P). (11.9.26)
1-e 6
k=1

The constant term is absent by virtue of the first equations in (12). As we


will see below the coefficients in front of the cosines will be not used in the
further calculation provided that only secular terms are considered. The
coefficients of sines are equal to

(11.9.27)

We can apply the equality

'P

2 arctan ~ --e
- tan -cp = V~J
l+e 2
1 - e2 d'ljJ 'ljJ ,
l+ecos
o
630 11. Perturbation theory

to estimation of the integral

J J+
211"
~
27r
(ak sin k<p + bk cos k<p) 2 arctan 1 - e tan '£2 d<p
1 e
o

J
211" 211"
~J 1 + d'ljJ
ecos'ljJ
(ak sin k<p + bk cos k<p) d<p
o 'Ij;

1] _ bk~1
211"
ak [~~ 1211" cos k'ljJd'ljJ - sink'ljJd'ljJ .
k 27r 1 + ecos'ljJ k l+ecos'ljJ
o o
By means of substitution ei'lj; = a, the integral

h =
1
27r
1+
211"

1
eik'lj;d'ljJ
e cos 'ljJ
o

f f
is reduced to the form
2 akda 2 1 akda
Ik= 27ri ea 2 +2a+e = 27ri~ (a-ad(a-a2)'
where integration is carried out over a unit circle. The roots of the denom-
inator aI, a2 are given by

al = -~ (1-~), a2 = -~ (1+~),
iali < 1 and ia2i > 1. By Cauchy's theorem we obtain

h = 2d = (_l)k (1- yI-e2)k


e (al - a2) ek yI-e2
It is a real-valued number, hence

~
27r
1+
~
cosk'ljJd'ljJ = (_l)k
1 ecos'ljJ
(1- ~)
ek~
k

,
1
27r
1+
211"
sink'ljJd'ljJ
1 ecos'ljJ = O.
o o
(11.9.28)

With the help of eqs. (22) and (25) we arrive at the equation determining
the secular change in the time of pericentron passage

dto
n- = (11.9.29)
du
11.10 Variational equations 631

where the coefficients ak are given by eq. (27). Limiting our analysis to
terms of first order in e we find

n ~: = -Cl (1 + 3e) (1- 3sin2 wsin2 i) . (11.9.30)

This formula is suggested in [96J.

11.10 Variational equations


The differential equations of perturbed motion (2.4) obtained by the method
of parameter variation are quite accurate. When the auxiliary problem for
the Hamiltonian function Ho differs from the initial one in small terms,
then the new variables of these differential equations (they are constant
in the auxiliary problem) are slowly varying functions of time which sub-
stantiates the applicability of the technique of approximate integration. In
contrast to this, the forthcoming way of consideration of perturbed motion
is based on construction of approximate differential equations for presum-
ably small deviations of perturbed motion from a prescribed unperturbed
motion. Taking into account only the first order of these deviations we
reduce the problem to consideration of a system of linear differential equa-
tions referred to as the variational system. Its integration is simplified by
the possibility of obtaining certain particular solutions whose number coin-
cides with the number of arbitrary constants in the solution of the problem
of the unperturbed motion.
The system of differential equations of the first order

(11.10.1)

is considered. The assumption that the right hand sides of these equations
do not explicitly contain t is essential.
Let us consider a certain particular solution of system (1) containing
k ::::; n independent constants

(11.10.2)

which is the unperturbed motion. The initial values of variables qs in the


unperturbed motion are designated by f~, that is

(11.10.3)

The difference between the perturbed and unperturbed motions is caused


by, first, different initial conditions and, second, by additional terms on
the right hand sides of equations (1). One can say that the solutions are
close in some sense or another only under the assumption of smallness of
variations. For instance, if we know the motion of a shell near the surface
632 11. Perturbation theory

of non-rotating earth and some prescribed initial conditions we can speak


about a perturbed motion due to the earth rotation and small deviations
in initial conditions. The theory of perturbed motion should, for example,
indicate a means of correction of the expression for the range obtained from
the solution of the unperturbed motion. However one can not expect that
the theory will deliver a satisfactory result when the resisting force of the
air is taken into account and the solution for the unperturbed motion in a
vacuum is at our disposal.
Along with system (1) let us introduce into consideration the following
system of differential equations

(11.10.4)

with fL being a small parameter. The perturbed solution is the solution of


this system satisfying the initial conditions

at t = to qs = f~ + x~, (11.10.5)

where f2 and x~ denote the initial values (3) of the unperturbed motion
and the small values termed the initial perturbations, respectively.
Instead of qs we introduce the new variables - perturbations Xs - by
means of the formulae

qs=fs(t-to,CI , ... ,Ck)+X s (8=1, ... ,n). (11.10.6)


Differential equations (4) are rewritten in the form

js + Xs = Qs (h + Xl,··· , fn + Xn) + fLips (h + Xl,··· , fn + Xn; t).


(11.10.7)

Functions fs are solutions of system (1), i.e.

(11.10.8)

so that

Xs = Qs (h + Xl,··· fn + Xn) - Qs (h,··· , fn) +


,

fLips (h + Xl,·" , fn + Xn) (8 = 1, ... , n). (11.10.9)


This system of differential equations, together with the initial conditions

at t = to (11.10.10)
determines the perturbed motion. Of course, eqs. (9) and (10) are simply
another form of the problem (4), (5). The next step is to use a presumed
smallness of perturbation and replace the exact equations (9) by their ex-
pansions in series in terms of powers of X at least for a finite time interval

to:::; t < to + T. (11.10.11)


11.10 Variational equations 633

Then we arrive at the system of equations

(s=I, ... ,n),

(11.10.12)

where the zero subscript implies that expressions (2) for variables qs in
the unperturbed motion should be substituted into the value in question.
Functions Rs contain powers of perturbation Xs of order higher than first
and the term

fJL
n (8<P
a s)
X r· (11.10.13)
r=1 qr a
This can be deemed as being a small value of higher order since it is mul-
tiplied by fJ. The first-order approximation is thus the following system of
inhomogeneous linear equations

xs=t(~Qs) xr+fJ<Ps(!I, ... ,fn;t) (s=I, ... ,n) (11.10.14)


r=1 qr a

with coefficients depending explicitly on time. This system is referred to as


the variational equations.
The feasibility of the above construction is substantiated by the theo-
rem on continuous dependence of solutions of the system of differential
equations on parameters and initial values: given a system of differential
equations

qs = Qs (q1, ... , qn; t; fJ1, ... , fJ p ) (s = 1, ... , n) , (11.10.15)

whose right hand sides are continuous with respect to all arguments qr, t, fJi
and have continuous partial derivatives with respect to these arguments in
a domain D. Then the system integrals, which are equal to q~ at t = to, are
continuous functions of to, q~, fJi and have continuous partial derivatives
with respect to to, q~, fJi in a sufficiently small domain 8, see [33].
The theorem should be understood as follows: let

qs (t, to, f 1,···,


a fa. a ... , fJ 0)
n' fJ1, p (s=I, ... ,n) (11.10.16)

be the solution of the system of differential equations (15) for fixed values
fJ? of parameters fJi and let qs = f~ at t = to. Then, for sufficiently small
positive values Cs one can determine positive values 'T]s and Mi such that,
for

Iq~ - f~ I < 'T] s' IfJi - fJ? I < Mi (s = 1, . .. , n; i = 1, . .. , p) ,


(11.10.17)
634 11. Perturbation theory

the solution of the above problem

qs (t,to,q~, ... ,q~;lll"" ,Ilp) (s = 1, ... ,n) (11.10.18)

in a certain time interval (11) will satisfy the inequalities

iqs(t,to,qL··· ,q~;lll"" ,Ilp)-qs(t,to,fp, ... ,f~;Il~,'" ,Il~)i <cs


(s = 1, ... , n). (11.10.19)
It is important to mention that values of'rJs and Mi determined through
Cs depend, in general, on T. The greater T is the smaller these values
are. If'rJs and Mi can be chosen in such a way that inequalities (19) hold
however great value t > to may be, the unperturbed motion (16) will be
stable in the sense of Lyapunov. In passing we notice that Lyapunov himself
studied dependence only on initial conditions. It is essential that stability or
instability in the sense of Lyapunov can be stated by means of the analysis
of variational equations (except for the so called special cases).
The questions of the stability of motion are beyond the scope of the
present book. We are content with the possibility of considering perturbed
motion on a finite time interval. From this perspective it is would be ad-
missible to consider for example the problem of stability of the pendulum
in the vicinity of its upper equilibrium position. The solution (16) will be

q = 0, q = 0,
where q is measured from the vertical position. Taking, for instance, c = 10
and E = 10 / sec we can determine such small values of the initial devia-
tion Qo and initial velocity qo that within the time interval T = 1 hour
inequalities (19)

hold and look for the solution of equations for perturbed motion for these
initial values. Of course this example [62] indicates only the necessity of a
reasonable choice of problems studied by the method of variational equa-
tions and, on the other hand, illustrates the fundamental significance of the
concept of Lyapunov stability.

11.11 On integration of variational equations


Let us consider the system of linear homogeneous equations

.
Xs = ~
r=l
(8
~ 8Qs )
qr 0
Xr (s=I, ... ,n), (11.11.1)
11.11 On integration of variational equations 635

corresponding to the variational system of equations (10.14). The next


remark essentially simplifies the search of the particular solutions and even
the complete integral of system (1) in some cases: the partial derivatives
of each of the integration constants C i of solution (10.2) of the system of
equations for the unperturbed motion (10.1)

(i) _ 81s
7]s -8Ci (s=I, ... ,n; i=I, ... ,k), (11.11.2)

as well as its time-derivative

(11.11.3)

are partial solutions of the system of linear homogeneous equations (1).


In order to prove this, we differentiate identities (10.8) with respect to
C i and t

.. ~8Qs'
Is == ~ 81r Ir.
r=1

This can be recast in the following form

'(i)=~(8Qs)
7]s - ~ 8 7]r(i) , '(HI)
7]s = ~ Qs (8
8 ) 7]r(k+l) . (11.11.4)
° °
~
r=l qr r=l qr
We arrived at eq. (1) which confirms the above said. Notice that for k = n
the particular solution (3) belongs to the set of solutions (2), since it can be
obtained by differentiating with respect to to, which is one of the integration
constants of solution (10.2).
The solution of Cauchy's problem for the system of equations for the
unperturbed motion (10.1)

qs=ls(t-to;q~, ... ,q~), (11.11.5)

is assumed to be given, that is

°
qs = Is (0 0)
0, ql , ... , qn' (~ql?s. )t=to = u.',s,
U • (11.11.6)

{jis being the Kronecker delta. From relationships (2) we obtain the com-
plete system of solutions of the homogeneous system (1)

'n(i)
'Is
(t) = 8q?
81s (S, z. = 1, ... ,n) . (11.11.7)

It satisfies the initial conditions at t = 0

7]~i) (0) = {jis (s,i = 1, ... ,n), (11.11.8)


636 11. Perturbation theory

forming a unit matrix. It is known, e.g. [82], that the Wronskian of the
system of linear differential equations is equal to

D(0) exp (Jo t (~Qs) dt)


(1) (n)
D(t) =
"'1 "'1
=
t
(11.11.9)
(n) (n) s=1 qs 0
"11 ",n

and is non-zero in as much as D (0) = 1. Solutions (7) form the fundamental


system which is a system of independent solutions with the unit matrix of
initial conditions for the homogeneous system of differential equations (1).
For this reason
n
Xs = L x~",~k) (t) (8 = 1, ... ,n) (11.11.10)
k=1
is the solution of this system.
The solution of Cauchy's problem for the variational system of equations
(10.14)

(11.11.11)

can be readily found by means of the method of parameter variation. We


take
n
Xs =L Ck (t) ",~k) (t) , (11.11.12)
k=1
with C k being determined from the following system of equations
n
'"' •
~ Ck (t)",s(k) (t) -_ /LiPs (It, ... ,In,. t) . (11.11.13)
k=1

It, (::)
It follows from eqs. (13) and (9) that

6, (t) ~ Mexp ( - 0 dt) t.<I\.~' (t), (11.11.14)

where llrk (t) denotes the algebraic adjunct ofthe element of the r-th row
and the k - th column of determinant (9). By virtue of eq. (12) we obtain
the solution of Cauchy's problem for the variational system of differential

I It, (::)
equations (11)

x, ~ t, "I') (t) [X2 + t.<I\.~dt)exp (-


M 0 dt) 1'
(11.11.15)
11.12 Equations for perturbed motion of a particle 637

FIGURE 11.1.

11.12 Equations for perturbed motion of a particle


The equations for the unperturbed motion of a particle with unit mass are
taken in the form of the natural equations

(11.12.1)

Let us refer to the trajectory C of particle M under this motion as


supporting curve. Further, 'T , n, b denote the unit vectors of the natural
trihedron, p and e> stand for the curvature radius of curve C at point
M and the arc along the trajectory, respectively. The force acting on the
particle in the unperturbed motion is designated by F, see Fig. 11.1.
Along with the unperturbed motion we consider the perturbed motion.
The latter can be treated as motion of another particle M* subject to force
F*. The forces F and F *, as well as the initial conditions for the particles
M and M* differ slightly from each other and this allows us to consider
equations for the perturbed motion as variations from the supporting mo-
tion.
The position of point M on the supporting curve is determined by the po-
sition vector r (e» . Let the normal plane P of the supporting curve passing
through point M* intersect this curve at point Ml with the position vector
r (e> + c). Let 'Tl , n 1 , b 1 denote the unit vectors of the natural trihedron
of the supporting curve at point M 1 , whilst 1/ and (3 denote projections of
vector MIM~ lying in the plane P on directions nl , b 1 , respectively. Then
position vector r* of point M* can be cast in the form

(11.12.2)
638 11. Perturbation theory

Using Frenet's formulae (2.18.9) we obtain

and, thus within the accuracy of the squares of small values, we have

r* = r - /-p) + n
+ Tc (11 c2 - T
( 1/ + 2p c(3) +b Cl/) .
( (3 + T (11.12.3)

While calculating the vectors of velocity and acceleration of point M* we


restrict our consideration to terms of first order. Then we obtain

v* 1/&) + n (.1/ + -;;


= v + T ( c. - -;; c& - y
(3&) + b (.(3 + T
1/&) '

W
* [.. v& 1/&2 I
=W+T c--;;+y p --;;--p
I/O- & (. c& (3&)]
I/+-;;-y +
. E& c&2 I cO- iJ& (3&2 I
n [I/+---P + - - - + - T
(30-
--+
p p2 P T T2 T
& ( E--;;
P 1/&) -T
& (.(3+T
1/&) ] +

.. v&
b [ (3 + - - 1/&2
- TI + -I/O- + -T& (.1/ + -c& - -(3&)]
T T2 T p T '
(11.12.4)

where a prime denotes differentiation with respect to arc length a. Casting


the vectorial equation of motion as follows

W* -w = F* - F,

we obtain three variational equations

.. vCr 1/ 1/ p' c (3 *
c - 2 - - - F + - F --F +-F =F -F
P P r P n P n T n r Tl
(11.12.5)

.. ECr c cp' 1/
1/ + 2- + -Fr - -Fn - -Fn-
p p p p
aiJ (3
2y - TFr + pFn
T2
( I
(3T - 1/
)
=
*
Fn - Fn, (11.12.6)

(11.12.7)
11.12 Equations for perturbed motion of a particle 639

In these equations iT and &2 are replaced by expressions from the equa-
tions for the unperturbed motion (1). As follows from the derivation, r;, F;;
and F; denote the projections of force F* acting on point M* in the per-
turbed motion on axes of the natural trihedron at point M. Let us calculate
these projections in detail.
It follows from expression (2) for vector r* describing the position of
point M* that the values
(11.12.8)
can be viewed as curvilinear coordinates. The unperturbed motion cor-
responds to the values = a, = q6 q5 q3
= 0. The basis vectors of this
coordinate system are determined by formulae (B.4.4)

r1 = ~:~ = 7 (1 - ~) - n1 ~ + ;1 b1, r2 = n1, r3 = b1.


Then, applying Frenet's formulae and neglecting the terms of second order
we find

r1 = 7 (1 - ~) + n (~ - :) + b~,
E E
r2 = -7-
p
+ n + b-
T'
(11.12.9)
E
r3 = -nT + b.
These formulae can be obtained by differentiating expression (3) with re-
spect to E, // and (3.
The covariant components of the metric tensor are
2// (3 //
911 = 1 - p' 912 = - T' 913 = T'
922 = 1, 923 = 0,
933 = 1.
They are calculated with the first order terms. With the same accuracy the
covariant components obtained by means of eq. (B.1.11) are given by
2// (3 //
9 11 = 1 + p' 9 12 = T' 9 13 = - T'
9 22 = 1, 9 23 = 0,
9 33 = 1.
Using this result and formulae (B.1.14) we obtain vectors of the co-basis

r1 =7 (1 + ~) + n ~,
r2=7(~-~) +n+b f , (11.12.10)

// E
r3 = --7 - -n + b.
T T
640 11. Perturbation theory

The vector F* is represented in terms of the covariant components F;


by expression (B.2.4)

(11.12.11)

In the case of potential forces

F* . fj r * = F * . rs fj q = - fjll = -~
S oIl fj q
uqs
S
,

the potential energy II being considered as a prescribed function of the


curvilinear coordinates. We obtain

F* = _ oIl (11.12.12)
S oqS'
For the unperturbed motion

(F*)o = F = FTT + Fnn = - (~~ r s ) 0

- [( ~~ ) 0T + (~~) 0 n + (~~) 0 b]
and thus

_ olIo _ F
oa - To -
(OIl) F
OV 0 = n, (11.12.13)

where 110 = II (ql, q6, q3) = II (a, 0, 0) implies the potential energy on the
supporting curve. Returning to expression (12) for force we have

orr
oqS =
( oqS
oIl ) 0 + oqk2oqS(0
11) 0 (k
q - qok) + ...

or, taking into account eq. (13),

(11.12.14)
11.12 Equations for perturbed motion of a particle 641

Due to eqs. (13) and (14) we have

These expressions should be inserted into the right hand sides of eqs. (5)-
(7). Such equations in the problem of focusing electron beams are consid-
ered in [32].
Thus, constructing variational equations requires knowledge of forces and
the second derivatives of the potential energy

(8
uv
IT)2
>l 2 0 '

both being calculated on the supporting trajectory.


Let us also consider the case of the resisting force depending upon the
velocity

F* = -f(v*)r*. (11.12.16)

By virtue of eq. (4)

. va
v * =0-+10--,
. }
r* =~v* =: + (v + lOa _ (3a) ~ +
v*
n
p T 0-
b (fi + av)
T
~.
0-
(11.12.17)

In the unperturbed motion

F =-f (v) r, Fr =-f (v), Fn = 0, Fb = o. (11.12.18)

Noticing that

f (v*) = f (a) + l' (a) (E _ v:) ,


642 11. Perturbation theory

we obtain

F; - FT = -1' (a) (t _1/:) ,


F*
nap
= _ f (a) (v + Ea _ j3a) T '
(11.12.19)

F; = - f c:) (iJ + a;) .

11.13 Perturbed Keplerian motion over a circular


orbit
In the central force field
r
F = -'ljJ (r)- (11.13.1)
r
the particle under certain initial conditions can move along a circular orbit.
On this orbit
r
n=-- (11.13.2)
r
and thus

(11.13.3)

The second equation shows that the velocity of the particle retains a con-
stant value throughout this motion. Due to the first equation this value
is

v = vr'lj; (r). (11.13.4)

Thus, the circular motion occurs when the particle gains the initial velocity
which is perpendicular to the radius vector and whose value is given by eq.
(4). The radius vector r* of the particle is given by expression (12.3) in
which we should set p = r, T = 00. We obtain

r* = r + T(E - E;) + n (1/ + ~:) + bj3. (11.13.5)

The potential energy of the particle under the perturbed motion in the
central field equals

J
T'

II (r*) = 'lj; (x) dx. (11.13.6)


11.13 Perturbed Keplerian motion over a circular orbit 643

The second derivatives of this expression are required for formulae (12.15).
To this aim we notice that
or* ~r* . or* = ~r* . (-~'T + n) ~ -~ + .!!.-,
Oll r* Oll r* r r* r*
or* ~r*. or* = ~r*. b = ii.
0(3 r* 0(3 r* r*

Hence,

oll
-='I/J (r *) -~--'I/J
or* r (*)
r +-'I/J
1I ()
r,
Oll Oll r* r*

and

( 02~)
Oll
= ['I/J (r) _ 'I/J' (r)] (or*) + 'I/J (r) = 'I/J' (r), }
r Oll r

( 021I)
0 0

= 0, (02~) = 'I/J (r). (11.13.7)


01l0(3 0 0(3 0 r

Taking into account eqs. (12.15), (3) and (7) we cast the variational
equations (12.5) as follows

(11.13.8)

In particular, in the problem of perturbation of the circular orbit of the


earth satellite
P, v2 , 2p, 2v 2
'I/J(r) = 2 =-, 'I/J (r) = -r3
- = --
r2
,
r r

and equations (8) take the form


v
E; - 2-v = 0,
r
v 3v 2
ii + 2-E; - - l l =0 (11.13.9)
r r2 '
.. v2
(3+2(3=0.
r
or
d2 11 dE
- +2- -311 =0 (11.13.10)
du 2 du '
644 11. Perturbation theory

Here we introduced the new independent variable


v
u = -t, (11.13.11)
r
which is the angle of rotation of the radius vector measured from the as-
cending node of the circular orbit. The solution of the system of equations
(9) is

~
du = 3 (2vo - E~)
v = 4vo - 2E~ + (2E~ -
+ 2 (2E~ - 3vo) cos u
3vo) cos u
+ 2v~ sin u,
+ v~ sin u,
} (11.13.12)

{3 = {30 cos u + {3~ sin u.

The first equation determines the perturbation of the orbital velocity whilst
the second and the third ones determine deviations from the supporting
orbital plane.
In order to take into account that the gravitational field is not central,
because of the geoid oblateness, it is necessary to add the forces (9.4) to
the right hand sides of equations (9). For this purpose, as explained in Sec.
11.10, it suffices to find their values on the supporting trajectory

<l'>T = - a -3-
R6 v2 sm
. 2· .
z sm 2u,

(3 3 )
r
2 2
<l'>n = aR+v 1 - - sin 2 i +- sin 2 i cos 2u , (11.13.13)
r 2 2
Ai.
'l'b = -aR6-3- sm
. 2' .
v2
z sm u.
r
Here i denotes the angle of the orbital plane with the equatorial plane of
the earth, a is a small parameter referred to as the constant of the earth
shape.
Introducing the new independent variables u and the non-dimensional
deviations of the circular orbit

al =a-,
Ro (11.13.14)
r
Ro being the earth radius, we obtain the following system of equations

(11.13.15)
11.13 Perturbed Keplerian motion over a circular orbit 645

The perturbed motion is a superposition of the motion (12) describing the


effect of initial perturbations on the motion caused by the perturbed forces.
In the plane of the supporting orbit these forces have a constant component
and the components with the double frequency of the satellite revolution
on the unperturbed orbit. The perturbation force has the frequency of
the satellite revolution on the orbit, that is, there exist the conditions for
resonance.
The solution of the system of equations (14) subject to zero initial con-
ditions has the form

15 1 = 20;1 [ (1 - ~ sin 2i) u - (1 - ~ sin 2i) sin u + 254 sin2i sin 2U] ,
"
U2 = -0;1 [(1 - "3
1. SIn2.)
z cos u - 6 . 2·z cos 2u - 1 + 21.
1 sm SIn2.J
z ,
1
15 3 = 20;1 sin 2i (u cos u - sin u) .
(11.13.16)

The radius vector of the particle in the perturbed motion is given by

1*
-r
r
= -n (Ro)
1 - -15
r
+ -Ro (715
2
r
1 + b153) , (11.13.17)

which follows from eqs. (2), (14) and (12.3). The corresponding unit vector
e; is as follows

(11.13.18)

Let m, m 1, k denote the unit vectors of the line of nodes for the un-
perturbed orbit, the perpendicular to it in the equatorial plane and the
perpendicular to this plane, respectively. Then, as Figs. l1.2a and 11.2b
show

7 = -msinu + cosu(m1 cosi + ksini) , }


n = -mcosu - sin u (m1 cosi + ksin i), (11.13.19)
b = -m1 sini + kcosi

and the expression for e; takes the form

e; m (cos u- ~o (h sin u) +
.. u
m1 ( coszsm Ro "
+ ~U1 .
coszcosu - Ro" .
~U3 sm z.) +

k (sinisinu + ~o 15 1 cosusini + ~o 153 cos i) . (11.1.3.20)


646 11. Perturbation theory

, k

{ a)

FIGURE 11.2.

Let the satellite pass the ascending node at u = 0 and let us denote the
value of argument u corresponding to the next passage as 21f + 1/J where 1/J
is small. In this position

cos () = e r* · k = sm
. .~ sm . ~.+ -Roo 3 cos ~. = 0.
. u + -Roo 1 cos u sm
r r
Then we obtain

21f . ~. +
1/J sm -:;:- 2a
(Ro) [2 (1 4"3.sm2.) sm.. + sm.. cos2.J = 0
- ~ ~ ~ ~

and moreover
'lj;
21f = - (
Ro
-:;:- ) 2
a (
3-
5 .
2 sm
2.
~. ) (11.13.21)

For this value of u the expression for e; is

e; = m+21fml { ! cosi + (~o ) a [COSi (2 - ~ sin


2 2 i) - sin 2 i cos i] }
or

e; = m _ 21fml (~o )2 acosi. (11.13.22)

Denoting the angular displacement of the node during one revolution of


the satellite by X we obtain

*
e r . m = cos
1f
(2" - X) ~ X= - 21f ( -:;:- )
R2o .
a cos ~. (11.13.23)
11.13 Perturbed Keplerian motion over a circular orbit 647

This coincides exactly with formula (9.14) for e = O.


Let us introduce the variable x defined as follows

(11.13.24)

It follows from the above said that cos (), being a periodic function of u
with period 27f + 7jJ, is 27f-periodic function of x. Turning to eq. (20) we
obtain

cos () = e r* · k = SIn . .Z + V3
. x + -x7jJ cos x ) + -Ro (s:vI cos U sm
. .Z ( sIn s: cos Z.) ,
27f ,

where we should replace u by x in the latter term since the second order
terms will be neglected. With the help of formulae (21) and (16) we find

cos () = sin i sin x { 1 - :f a [ (2 - ~ sin2 i) cos x - ~ cos2 x sin2 i + cos 2 i] }


(11.13.25)

Denoting the longitude of the satellite by A we have

x7jJ. Ro .
sin () cos A = e; . m = cos x - - sm x - - 81 (x) sm x
. ( .27f x7jJ ' )
sin () sin A = e; . ml = cosz smx + 27f cos x +

,
Ro [8 1 (x) cos x sin i - 83 (x) sin i]

or using the above formulae

sin () cos A = cos x { 1 + :f a [x tan x cos 2 i+

2- ~3 sin 2
i 5
sin 2 x - - sin 2 i sin 2 x
1}
cos x 6

:f
(11.13.26)

sin () sin A = cos i sin x { 1 - a [x cot x+

(2- ~ sin2 i) cos x - ~ sin 2 i cos 2 X - sin 2 i] }

and therefore
R2
tan A cositanx { 1- ,ga[x(cotx+tanxcos 2 i)+

2 - -2.
3
sm 2·Z 11
_ _ sin 2 i 1} . (11.13.27)
cos x 6
648 11. Perturbation theory

Now we express variable x in terms of time t. Since longitude). does not


appear in the expression for the potential energy of the perturbed forces,
there exists the integral of the moment of momentum about axis O(
2 2·
p),,=r* sin (}).=/3)".

for the perturbed motion. However the perturbations 8 1 ,82 ,83 and their
derivatives are equal to zero at the initial time instant which allows calcu-
lation of the constant (3)" under motion on the circular orbit with velocity
v
(3)" = (r x 'Tv) . k = -rv (n x 'T) . k = rvcosi.

Thus,
>. = rv cos i
(11.13.28)
r*2 sin2 ().
On the other hand, by differentiating expression (27) we obtain

co~i x
cos x
{I - R~r
a [cos 2 x + sin 2 xcos 2 i + 2xtanx cos 2 i+

2 - -2.sm 2 z. 11 ]}
3 (1 + sin 2 x) - -6 sin 2 i .
cos x

Expressing x, replacing>' and sin () cos). by means of eqs. (28) and (30) and
taking into account, that due to eq. (17),

r*2 = r2 { 1 + 2., a [ ( 1 - ~ sin2 i) cos x - ~ sin2 i cos 2x - 1+ ~ sin2 i] } ,


(11.13.29)
we obtain, after a little algebra, a very simple result

.
X
v
= ;: [1 + R~ (3 25. 2.)]
~a z . - sm (11.13.30)

From this equation we obtain

r
t - to =;x ~ (3- 2
[ 1 - -:;:2a 5.
sm z 2.)] , (11.13.31)

where to denotes the time instant of passage of the ascending node. The
next passage of the ascending node (in its new position) will be at x = 27r.
Designating this time interval by T we have

T = To [1 - . , a (3 - % sin2 i) ] , (11.13.32)
11.14 Equations for perturbed motion of a material system 649

with To being the period of the unperturbed motion on the circular orbit
r
To = 2n-.
v
In passing we note that another formula for period T was derived in [9].

11.14 Equations for perturbed motion of a


material system
In Sec. 7.8 we give a geometric interpretation of the motion of a holonomic
system of particles subject to stationary constraints as motion of the rep-
resentative point on Riemannian manifold Rn determined by the square of
the arc element

(11.14.1)
The generalised coordinates in the unperturbed motion qS are assumed
to be prescribed functions of time
qS = qS (t) (s = 1, ... ,n). (11.14.2)
These relationships compose a particular solution of the differential equa-
tions of motion which can be set in the form (7.8.12)

it" + { ;, } i/q"f = Qa = aa{3Q{3, (11.14.3)

expressing equality of the contravariant components of acceleration and


force of the representative point. Equations (2) can be considered as para-
metric equations of the trajectory of the representative point, i.e. the sup-
porting trajectory. The unit vector T of the tangent to this trajectory is
determined by its contravariant components
dqa 1.a
Ta - _ --q (11.14.4)
- deY - iJ
The generalised coordinates of the perturbed motion are denoted as qa +x a ,
where the perturbations x a can be deemed as contravariant components of
the perturbation vector
p = xar a , (11.14.5)

r a being the basis vectors.


The equations for the perturbed motion are obtained by replacing the
generalised coordinates qa in eq. (3) by qCi + XCi such that

(11.14.6)
650 11. Perturbation theory

where Christoffel's symbols and forces are taken for the perturbed motion.
Here /L is small parameter and thus the additional forces given by their
contravariant components cpa are included in the variational equations of
the supporting trajectory. Up to the first power in perturbations we have

{.~ L~ ~; }+.x;gt { ;, }, } (11.14.7)


Q* - qo. + X oqf3 + x oii'
with the forces being assumed to depend on the generalised coordinates
and velocities.
Finally we arrive at the variational system of equations

"0.+ X o'f3"'t~{
X
a
q q oqO (3"( }+2{ a }'f3''Y=
(3"( qx xf3oQo.+'f3oQa+ iF.o.
oqf3 x oqf3 /L'*',
(11.14.8)
which can be readily used for solving some problems. Of interest is the
transformation suggested by Synge [86] which reveals the tensorial charac-
ter of the quantities in eq. (8).
The vectors of velocity and acceleration are introduced into considera-
tion. Differentiating eq. (5) using rule (B.6.1) yields

(11.14.9)

and moreover

p = ro. [xo. + { ;"( } ilx'Y + { (3~ } qf3:t'Y+

qf3 qU x'Y o~u { (3~ } + (:to + { ;"( } qf3 x'Y) { 8: }


qu] .

Replacing here il by means of equations (3) for the unperturbed motion


we have

The coefficients of rain the expressions for p and p are the contravariant
*0 **0
components of these vectors. Let us denote them by x and x ,then

;;0. =:to. + { (3"(


a } ilx'Y = (dXdao. + { a } rf3 x'Y)
(3"(
a, (11.14.11)
11.14 Equations for perturbed motion of a material system 651

~Q = iQ {;8 }+ { (Ja, } { %8 }-
+ q~ q~ X O (a~~

{:8 }{%, }) + 2{f3a, } q~j;~ + { ;, } Q~x~. (11.14.12)

*a **0:
Now we replace j;Q and iQ in eq. (8) by x and x . The principal signifi-
cance of this transformation is that we use the components of the invariant
quantities, namely the vectors of velocity and acceleration. If we assume
<l>Q = 0 and the forces to depend only on coordinates, then we come to the
following equations

The expression

(11.14.14)

is the covariant derivative of the contravariant components of the force


vector Q. Recalling definition (B.14.2) of the Riemann-Christoffel tensor

a~O { ,~ }- a~~ { 8~ } +
{ ;f3 } { 8: }-{,: }{M} (11.14.15)

we can rewrite variational equations in the form

(11.14.16)

Using Ricci's theorem, see Sec. B.5, we can cast these equations in the
covariant components

(11.14.17)

The covariant components of the Riemann-Christoffel tensor are deter-


mined by formulae (B.14.5). They have the following properties:
1) symmetry with respect to pairs of subscripts ,8 and f3(J

(11.14.18)

2) skew-symmetry with respect to subscripts, and 8, f3 and (J


(11.14.19)
652 11. Perturbation theory

3) Ricci's identity

(11.14.20)

expressing the property of cyclic symmetry with respect to the three co-
variant subscripts.
In the case of potential forces equations (17) take the form

} :;).
(11.14.21)

The energy integral for the unperturbed motion is as follows

T + II = h,

whereas that for the perturbed motion is

where oh denotes variation of the potential energy. In the latter expression

Here we used the representation for the derivatives of aa{3 in terms of the
Christoffel's brackets derived in Sec. B.4. Taking into account eq. (11) we
obtain

and the relationship expressing the variation of the energy integral takes
the form

(11.14.22)

11.15 Systems with two degrees of freedom


Let us consider the holonomic conservative systems with two and three
degrees of freedom in detail. Generalisation to more complex systems would
require further knowledge of Riemannian geometry.
11.15 Systems with two degrees of freedom 653

This section is concerned with the systems with two degrees of freedom.
It is necessary to remember that the initial data are, first, the components
of the metric tensor aOl.{3 and, second, the equations for the unperturbed
motion (14.2) for n = 2. Basis vector rOi. having direction of the tangent
to the coordinate line along which parameter qOl. changes is no longer the
derivative of vector r belonging to the manifold with respect to qOl.' Vector
r can be determined only on an Euclidean manifold which includes the
Riemannian one under consideration.
The unit vector T of the tangent is determined in terms of its contravari-
ant components with the help of these data and formulae (14.4). The unit
vector of the first normal n in the Riemannian geometry is introduced by
means of Frenet's first formula which is just another form of eq. (7.5.22)

+ = wn, w = kir, (11.15.1)

where k denotes the first curvature. Frenet's second formula (for n = 2)

n= -WT (11.15.2)

is easy to obtain by differentiating the relationship T . n = 0 and taking


into account that n has direction T (or opposite) so that n· n = O. It is
worthwhile noticing that eqs. (1) and (2) are equivalent to relationships of
the form
*0 *0:
= wnOl., = = wnOl., =
*(}: *(}:
T n -WTOI.; T n -WTOI. (11.15.3)

linking covariant and contravariant components of vectors +, nand n, T.


Similar to eq. (14.11) the asterisk defines the following operation

;.01. = TOI. + { ;"1 }irT{3T'Y, iiOi. = nOi. + { {3a"1 } irT{3n'Y. (11.15.4)

Let us now represent the perturbation vector in the form

p=C:T+vn. (11.15.5)

Then, by virtue of eqs. (1) and (2) we obtain

jJ (t-WV)T+(v+wc:)n,
P (E - 2vw - vw - c:w 2 ) T + (ii + 2tw + c:w - vw 2 ) n.

From these equalities we obtain relationships of the sort

= P.T = r = aOl.a
. • *0' U *0 (T
c: - wv 01. x 'T r 01. x T

and arrive at four formulae

(11.15.6)
654 11. Perturbation theory

a oa **0
X
a _.. . .
T - E - 2vw - vw - EW ,
2}
(11.15.7)
a oa X
**0
n
a
= V.. + 2·EW + EW. - 2
VW .

Returning to the variational equations (14.21) and taking into account that

(11.15.8)

we come to another form of these equations

€- 2vw - vw - EW 2 + &2 RryOfJa (ET'YTOTfJna + VT'YnOTfJna) +


(ETO + vn O) TaVoIIa = 0, (11.15.9)

i/ + 2t-w + EW - VW 2 + &2 RryofJa (ET'YTOTfJTa + VT'YToTfJTa)


+ (ETO + vnO) naVoIIa = 0 (11.15.10)

In order to avoid a similar derivation for the case n = 3, we consider the


following invariant

(11.15.11)

determined by the two chosen vectors a and c.


To calculate this invariant we enter Ricci's tensor which is the symmetric
tensor of second rank defined as follows

A>'JL =~ EA'Y OEJLfJ a Rryof3a, (11.15.12)

where EA'Y O denotes the Levi-Civita symbol introduced by eqs. (B.1.16) and
(B.1.17). Using their properties and taking into account eqs. (14.18) and
(14.19) we can easily obtain the formulae

12 1 13 1
A = ~R2331' A = ~R2312'

A
22 = ~R3131'
1
A
23 _ 1
- ~R3112' (11.15.13)
33 1
A = ~R1212'

where lal = laofJl. These formulae contain all six independent components
of the Riemann-Christoffel tensor for n = 3. As follows from eq. (B.14.5),
for n = 2, there is only one independent variable which is

A 33 = R1212 2 = K. (11.15.14)
alla22 - a 12

In the theory of surfaces K means the Gaussian curvature.


11.15 Systems with two degrees of freedom 655

By means of eq. (12) we can obtain the expression for the Riemann-
Christoffel tensor in terms of Ricci's tensor

Rry8{3u =EA'Y8EI-'{3u AAI-'. (11.15.15)

To prove this we make use of the following identity

E,A'Y <E AXP = 8'xY 88P - 88x 8"Y'


U
P (11.15.16)

where 8~ denotes Kronecker's symbol. Substituting for Ricci's tensor in eq.


(15) its expression (12) yields

4"1 EA'Y8 E
AXP EI-'{3u E I-'TW R XpTW -_ 4"1 (8X8P
'Y 8 -
8X8P)
8 "y
(8T 8W 8T
{3 u -
8W)
u (3
R XpTW·
(11.15.17)

Expanding the expression on the right hand side and using properties
(14.18) and (14.19) of the Riemann-Christoffel tensor we finally arrive at
the required equality (15).
The invariant (11) to be calculated is written in the form

K(a,c) = A>'I-' E>''Y 8 T'Ya 8 EI-'{3u T{3C u •

Then, referring to representation (B.2.7) for the vector products

7" X a = p, 7" X C = q, (11.15.18)

we obtain

K (a, c) = A>'I-'p>.qw (11.15.19)

From this formula it follows that

K (a, 7") = K (7", c) = K (7", 7") = o. (11.15.20)

Let

a= c = n,
where n denotes the unit vector of the normal to the curve. Then

(11.15.21)

where b = 7" X n designates the unit vector which we refer to as the second
normal. Its covariant components are equal to

(11.15.22)

By analogy

K (b, b) = A>'I-'n>.nl-" K (n, b) = -A>'I-'b>.nl-' = K (b, n). (11.15.23)


656 11. Perturbation theory

These formulae correspond to n = 3. For n = 2 only K (n, n) is introduced


into consideration since due to eq. (19)

K(T,n)=O, K(T,T)=O. (11.15.24)

The vectors T and n belonging to R2 have the components T1, T2 and


n 1 , n 2 respectively (let us recall that kn Q is determined as the covariant
derivative of TQ). Thus, unit vector b (which does not belong to R 2 ) has
the single component

(11.15.25)

which is equal to unity because it is equal to the area of the square con-
structed by the unit vectors T and n. Hence, for n = 2, due to eqs. (14)
and (21)

K (n, n) = R1212 2 = K. (11.15.26)


alla22 - a 12

Let us return to eqs. (9) and (10). By virtue of eqs. (24) and (26) they
have the form

where

(11.15.28)

The integral of energy, due to eqs. (14.22), (5) and (6), is given by

a (E: - vw) ~rr = oh.


+ (ETQ + vnQ) uqQ (11.15.29)

As follows from eq. (7.5.29) for the unperturbed motion

..
a =
orr
---T
a .
aw = ---n .
orr a
(11.15.30)
oqQ ' oqDi

Using this result and differentiating eq. (29) we obtain the first equation in
(27). Then having the energy integral we can remove it from consideration.
The time-derivatives of TQ and nQ are determined by means of formulae
~rr in
(3). The time-derivative of the covariant component of the force -
uqQ
the unperturbed motion is determined analogously by means of the rule of
the covariant differentiation.
11.16 Systems with three degrees of freedom 657

Using eq. (30) we cast the energy integral in the form

iJE: - ac = 2iJvw + 8h. (11.15.31)

Then, taking the derivative of the second equation (30) with respect to
time and using the first one we find

(J"W
.. . . = -;:;--WT
+ (J"W all '" - . li n "'IIli",
(J"T
uq'"
or

(11.15.32)

Taking into account eqs. (31), (32) and (1) we write eq. (27) as follows

jj + v (iJ 2 K + 3iJ 2 k 2 + nlin"'Illi<T) + 2k8h = O. (11.15.33)


It contains only v and the parameters which are known from the unper-
turbed motion. Under the assumption that v has been found, the problem
of determining c from the energy integral (31) reduces to quadrature

J J~; +
t t

c = 2iJ kvdt + iJ8h (J"


co ~,
0"0
(11.15.34)
o 0

where co and iJo denote the values of c and iJ at the time instant t = 0
when the initial perturbation first occurs.

11.16 Systems with three degrees of freedom


We introduce into consideration unit vectors of the tangent T, the first
normal n and the second normal b. As before, n is given by Frenet's formula
(15.1) and b = T X n, so that T, n, b are determined at any point of R3
and makes an orthonormal trihedron. The first Frenet's formula is now set
in the form

(11.16.1)

where k(1) denotes the first curvature given by eqs. (7.5.22) and (7.5.23).
The subscript 1 is placed in parentheses as it indicates the number of the
values rather than its covariant character. Differentiating the relationship
n . T = 0 we obtain

n· T = -;- . n = -W(l)
and taking into account that n· n = 0 we can write

(11.16.2)
658 11. Perturbation theory

Here factor W(2) is introduced. By virtue of eq. (2) it is equal to

n· b = W(2). (11.16.3)

Keeping in mind that

and applying eqs. (1) and (2) we obtain the third Frenet formula

(11.16.4)

Here W(2) = k(2)&, k(2) being called the second curvature of the curve in
R3 ·
By applying Frenet's formulae (1), (2) and (4), we can represent the
perturbation vector p as follows

p = ET + vn + ,6b (11.16.5)

and obtain, instead of eqs. (15.6) and (15.7)

(11.16.6)

Furthermore

(11.16.7)

Taking into account relationships (15.20) we can write differential equa-


tions (14.21) under notation (15.11) in the form

E: - w(1)V - 2w(1)zi - wl1)E - W(1)W(2),6 + (ETO + vno + ,6bO) TUlluo = 0,


(11.16.8)

i/ + W(1)E - W(2),6 + 2w(1)E: - 2W(2)i3 - (Wll) + Wl2)) v +


&2 [K (n, n) v +K (n, b),6] + (ETO + vno + ,6bO) nUlluo = 0, (11.16.9)

,6 + W(2)V + 2W(2)zi + W(1)W(2)E +


- 2
- W(2),6

&2 [K (n, b) v +K (b, b),6] + (ETO + vno + (3bO) bUlluo = O. (11.16.10)


11.16 Systems with three degrees of freedom 659

The energy integral (14.22) is

. (.E -
a W(1)V
) all (8
+ -;::;-- 8 (3 8) = uh.
ET + vn + b
, (11.16.11)
uqa

Equations for the unperturbed motion should be written as follows

.. all a
a---T (11.16.12)
- oqa '

Differentiating the second equation with respect to time we find with the
help of Frenet's formulae and the two remaining equations (12) that

..
aW(1) -
2 -;::;--T
all a W(l) + n aTua' II u8 = 0 , (11.16.13)
uqa

whilst the third equation in (12) after the covariant differentiating yields

(11.16.14)

Equations (12) allows the energy integral (1) to be expressed in the form

(11.16.15)

which does not differ from eq. (15.31).


Using now eqs. (15), (14) and (13) we can transform eqs. (9) and (10),
to get

+ v[a
V.. . 2K
ll ( , II ) + 3W(1)
2 2 + II 8un 8 n u]
- W(2)

+ (3 [0- 2 K (ll, b) - W(2) + Il8un8bU] - 2W(2)/3 + 2k(1)8h = 0, (11.16.16)

(3.. + (3 [ 0- 2 K (b, b) - + Il8ub 8 bU]


w(2)
2

+ v [K (ll, b) 0- 2 + W(2) + Il8un8bU] + 2W(2)z/ = O. (11.16.17)

Quantity E does not appear in these equations. It is determined by quadra-


ture from the energy integral (15) with the help of eq. (15.34). Equation (8)
is excluded from consideration since it can be obtained by differentiating
eq. (15).
Let us recall that the invariants K (ll, ll) and K (b, b) are quadratic
forms of na and ba , respectively, whereas K (ll, b) is a bilinear form of
these quantities. These forms are given by eqs. (15.21) and (15.23), and
their coefficients are the components of Ricci's tensor calculated by eq.
(15.13).
660 11. Perturbation theory

11.17 Stationary unperturbed motions


The unperturbed motion for which the positional generalised coordinates
and the cyclic generalised velocities retain constant values is referred to as
stationary.
Provided that the unperturbed motion is stationary all coefficients in eqs.
(16.16) and (16.17) are constant and moreover W(2) = O. These equations
take the form

it + P2V + qf3 - 2W(2)iJ + 2k(1)bh = 0, } (11.17.1)


/3 + P2f3 + qv + 2W(2)V = 0,

where

(11.17.2)

A8,B being the covariant components of Ricci's tensor. The integral of energy
(15.34) is

J
t
tbh
E = 2&k(1) vdt + ;;- + EO. (11.17.3)
a
Synge following Routh as well as Thomson and Tait referred to the un-
perturbed stationary motion as kinematically stable if the value of the per-
turbation vector

(11.17.4)

remain bounded for any t however great it is.


The particular solution of the system of homogeneous equations corre-
sponding to system (1) is sought in the form

(11.17.5)

We obtain the system of homogeneous equations

(PI - r2) N + (q - 2w(2)ir) B = 0, }


(11.17.6)
(q + 2w(2)ir) N + (P2 - r2) B = O.

The characteristic equation is

(PI - r2) (P2 - r2) - q2 - 4w(2)r2

r4 - (PI + P2 + 4W(2) ) r2 + PIP2 - q2 = O. (11.17.7)


11.18 Examples 661

Let us denote the roots of this biquadratic equation as r~ and r~. Pro-
vided that P1P2 - q2 = 0, eq. (7) has a double zero root and the solution of
the system of inhomogeneous equations (1) has terms proportional to t 2 . If
r~ (or r~) are negative or complex-valued then irl (or ir2) is either positive
or has a positive real part which means that the corresponding particular
solution grows without bound with t. Thus, positiveness of roots r~ and r~
of the biquadratic equations is the necessary condition of stability under
the above definition. The system is stable when the inequalities

Pl + P2 + 4W~2) > 0, (Pl + P2 + 4W~2)) 2 > 4 (P1P2 - q2) > 0 (11.17.8)

hold. This would be sufficient at oh = 0 for small perturbations for which


the total mechanical energy remains unaltered with accuracy up to first
order powers in these perturbations. Then the constant term in equations
(1), as well as the term proportional to t in eq. (3), vanish. However if
oh =I- 0 then the particular solution of system (1) is

11= (11.17.9)

and c remains finite under the following condition

P1P2 - q2 = 4W(l)q·
2 (11.17.10)
The frequencies of oscillations about the stationary motion are rl and
r2. It would be an error to think that inequalities (8) themselves ensure
bounded values of 11 and f3 because the variational equations (1) are derived
under the assumption that c is bounded.
A similar analysis of a simpler case of two degrees of freedom leads, at
oh = 0, to the inequality
(11.17.11)
which, after Synge, gives the stability condition for the stationary unper-
turbed motion. If oh =I- 0 the equality
(11.17.12)
should be added to inequality (11). The frequency of oscillation about the
stationary motion is equal to 2w.

11.18 Examples
11.18.1 Two particles attached together with a string
Two particles Ml and M2 of masses ml and m2, respectively, are attached
together with a weightless inextensible string of length t. The string passes
662 11. Perturbation theory

FIGURE 11.3.

through an opening 0 in a smooth horizontal table. Particle MI remains


on the table, the part OMI of the string is tight and the part OM2 moves
vertically along axis Oz, see Fig. 11.3.
Let us denote the polar coordinates of particle MI by T = OMI and 'P
then the kinetic and potential energies of the system are given by

T = ~ (mIr2 + mIT2rp2 + m2r2) , II = -m2g (l- T) = m2gT + const.


(11.18.1)
Constructing the differential equations of motion by means of Lagrange's
equations we obtain
(11.18.2)
In what follows we study the stability of the stationary motion
m2g
T = TO = --.-2 ' rp = rpo = const (11.18.3)
ml'PO
which is one of the particular solutions. Under this motion particle M2 does
not move whilst particle MI moves on a circle of radius TO with angular
velocity rpo. The string tension at any point is m2g = mlTorp5.
We notice that the motion of the system can be compared with the
motion of particle m = ml on the surface of a cone with the vertical axis,
with vertex 0 beneath and the half opening of the cone a defined by

cot a = - 2.
~
ml
(11.18.4)

Indeed, in cylindrical coordinates the surface of the cone is prescribed by


the equation
z = T cot a. (11.18.5)
11.18 Examples 663

Thus, the kinetic and potential energies of the particle are

T = ~m [.;.2 (1 + cot 2 a) + T2~2], II = mgz = mlgT cot a = m2g*T

which coincides with expression (1) provided that g* = gJm2/ml.


The motion of the representative point on Riemannian manifold R2 is
considered. The covariant components of the metric tensor on this manifold
are equal to

(11.18.6)

Equations of the unperturbed motion (14.2) have the form

ql = TO, q2 = ~ot. (11.18.7)

The squared velocity &2 under this motion is

a.22T
= 2·2
= mlTo<po, (11.18.8)

whereas the covariant components of the unit vector of the tangent are
given by
d 2 •
T2 -....!L _
<Po (11.18.9)
- da - &.

The vector of the tangent is a unit vector since, due to eq. (8),

We find Christoffel's symbols of the second kind by comparing eq. (2) with
equations of motion in the form of eq. (14.3). The only non-zero symbols
are

(11.18.10)

Using formulae (15.3) we obtain

that is, in our case

(11.18.11)

Value w = k& is determined from the condition that n is a unit vector


664 11. Perturbation theory

We find

1 1
n = - -y'rm=l=+=m=2 '
(11.18.12)

Calculation by formulae (15.26) and (B.14.5) yields K = 0 which we


might expect since the Gaussian curvature of the conical surface is equal
to zero. Expression II b ,6n b n,6 is zero as well because of

IIn =
(PII
or2 +
{I} all
11 or = O.

Equation (15.33) takes the form

i/ + 3w 2 v + 2 8h = O. (11.18.13)
rov'ml + m2
Stability of the unperturbed motion in the sense of Synge is possible only
for perturbations for which the total energy of the system is conserved (i.e.
8h = 0) as equality (17.12) does not hold. For the problem in question

2h = 2 (T + II) = (ml + m2) 1'2 + mlr2ifJ2 + 2m2gr,


thus, applying (3) we obtain

8h = (mlroifJ6 + m2g) 8r + mlr5ifJo8ifJ = ml ( 2roifJ68r + r5ifJo8ifJ) = ifJ o8f3<.p


since l' = 0 under the unperturbed motion. Here f3<.p = mlr5ifJ denotes the
resultant angular momentum about axis Oz and, for this reason, 8h = 0
for the perturbations which do not affect the angular momentum. Such a
perturbation can be realised in the case when particles Ml and M2 get
impulses directed along the string. In this case, the unperturbed motion is
stable under the above definition.
Turning our attention to eqs. (15.33) and (15.34) we adopt the following
initial conditions

at t=O Xl=x6=(8r)o,
(11.18.14)
X2 = x6 = (8cp)o,

where

(11.18.15)

It is necessary to find the initial values of v, V, c, E. By means of eqs. (15.5),


(12) and (9) we obtain
11.18 Examples 665

and thus

(11.18.16)

Next, due to eqs. (14.11) and (15.6) we have

xl + {212}T 2x 2a- = (it + WE) nl, x2+ C21}T2X l a- = (E - wv) T2,


which leads to

(11.18.17)

which, of course, can also be obtained by differentiating eq. (16). Condition


(15) together with (12) yields

Eo = 2vow. (11.18.18)

The general solution of equations (15.33) and (15.34) is written as follows

v = Vo cos V3wt + ;;w sin V3wt, }


(11.18.19)
E = Eo + ~ [vo sin V3wt + ;;w (1 - cos V3wt)] .

It can be easily proved that condition (18) is met. Perturbation of the polar
radius 8r and the angular velocity 80 are determined by eqs. (19), (17) and
(16)

8r = (8r)0 cos V3wt + 8;0 sin V3wt, }


v3w
r080 = -200 [( 8r)0 cos V3wt + ~: sin V3wt] = - 2008r.
(11.18.20)

The angular momentum of the system about axis 0 z is seen to be conserved


under the perturbed motion at any time instant. Clearly, expressions (20)
could be easily obtained directly by varying the integral of energy (under
the assumption that the total energy is conserved up to first order terms)
and taking into account the integrals of angular momentum under both
perturbed and unperturbed motions. The above was intended to illustrate
the calculations needed to analysis of mechanical problems in terms of
Riemannian geometry.

11.18.2 Stability of regular precession


Let us consider the problem of stability of the regular precession of a heavy
gyroscope which is a symmetric body rotating about a immovable point.
666 11. Perturbation theory

Describing the position of the axes fixed in the body by the three Euler's
angles {} = ql, 'l/J = q2, ifJ = q3 we come to the following expressions for the
kinetic and potential energies

T= ~ [A (~2 + ?j? sin2{}) + C (p + ~ cos {}) 2], II = Q Zc cos {}


(11.18.21 )

and thus
all = A, a12 = °
a22 = A sin 2 {} + C cos 2 fl,
a13 = 0,
a23 = C cos {},
}
(11.18.22)
a33 = C.

Resolving Lagrange's equations for 19,;P, 'P we have


.. C - A ·2 C . Qzc
{}+ -A-'l/J sin{}cos{}+ A0'l/Jsin{} = A sin {}, (11.18.23)

.. 2A- C· . C 0~
'l/J + A 'l/J{} cot {} - A sin {} = 0, (11.18.24)

'P - 1
sin {}
(A - C cos {} + 1
-A-
2 ) .. C·
'l/J{} + A 0{} cot {} = 0, (11.18.25)

and then

{I} = -A-
22
C-A
sin {} cos {},

{212} = 2AA- Ccot{},


{3} = - -1- (A-C
21 A 2sin{}
- - c o s 2 {}+ 1)
'

°°
Under stationary motion of the regular precession the angular velocities
of precession ~ and spin as well as the angle of nutation {} retain constant
values. Assuming ~ = ~o, = 00, {} = {}o we satisfy eqs. (24) and (25)
whereas eq. (23) determines the equality which should relate these values
in order to ensure regular precession
C - A'2 C'. Qzc
-A-'l/Jo cos {}o + A'l/JoifJo = A' (11.18.27)

Under the condition

(11.18.28)
11.18 Examples 667

eq. (27) has two real-valued roots

';;;0 = 2 (A _ ~) cos {}o [Cif?o =f JC if?6 -


2 4 (A - C) Qzc cos {}o] ,
(11.18.29)

defining slow and fast precessions.


For the perturbed motion

We write equations for the perturbed motion in form (14.8) where Xl


X, X2 = ~,x3 = 17. Taking into account that Christofell's symbols depend
only on ql = {} we find

{) .(3 "'I
X q q aq8
a {(3,1 } X1('2C-A . C
'l/Jo-A- cos2{}0 + if?0'I/J0 A cos{}o ) ,

8 .(3 "'I
X q q aq8
a {a}
(3, = 0 for a = 2,3,

since the non-zero Christoffel's symbols with superscripts 2 and 3 have


subscript 1 and {j = ql = 0 for the unperturbed motion.
We obtain the following equations

Here it is taken into account that

Q'" -_ a
"'IQ __ ",loll _
I - a a{} -
Qzc .
A sm u,
"0

since all = A-I,a I2 = a l3 = O.


The two remaining equations are

t + 2{221}';;;OX + 2{321}if?ox 0,

i?+2L31}';;;oX+2{331}if?ox 0

or

~ + 2{ 221 },;;;OX + 2{ 321 }if?OX a,

17 + 2{231}';;;OX + 2{331}if?ox (3,


668 11. Perturbation theory

where a and /3 are constants. They can be expressed in terms of variations


8/3'P and 8/31jJ of cyclic coordinates
8T 8T
/3'P = 80' /31jJ=- ..
8'l/J
These are zero if we assume that the angular momenta of the rigid body
about its symmetry axis and the vertical axis retain the same values as
under the unperturbed motion. This occurs in the case of perturbation of
the motion by an impact in the vertical plane passing through the gyroscope
axis.
Now substituting for Christoffel's symbols eq. (26) we arrive at the sys-
tem of equations

- A . C QZc
x + x ( 'l/J. o2 -CA )
- cos 219 0 + 00 'l/Jo A cos 190 - A cos 190 +
2 (C - A) . . C·. C..
A 'l/Jo~ sm 19 0 cos 19 0 + A 'l/Jo'TJ sm 19 0 + A i.{Jo~ sm 19 0 = 0,
2A-C. C)
F: +x ( A 'l/Jo cot 190 - A sin 190 00 = 0,

'TJ +x [ - sin119
0
(A- A-C 2
- cos 19 0
C. ]
+ 1) .'l/Jo + A i.{Jo cot 190 = o.
(11.18.30)

Inserting the obtained expressions ~ and 'TJ into the first equation in (30)
and removing the angular velocity of the spin 00 by means of eq. (27) we
obtain the differential equation

(11.18.31)

where
Qzc
J-L=-·-2· (11.18.32)
A'l/J o
Hence, the perturbed motion about the regular precession is a harmonic
oscillation of the angle of nutation with frequency

(11.18.33)
12
Variational principles in mechanics

12.1 Hamilton's action


The basic statements of dynamics, which are Newton's axioms and D'Alem-
bert's principle, allows us to formulate the laws of motion in the form of
differential equations of motion. However they do not exhaust all the ways
of representing the laws governing the motion of material bodies. An alter-
native is variational statements dealing with the stationary properties of
certain values and enabling the complete replacement of the above state-
ments.
One acts in the same way in geometry. An object may be prescribed
by its differential properties in terms of the differential equations or by a
certain variational requirement. For instance, a geodesic line is defined as a
curve on the surface whose principal normal coincides with the normal to
the surface. This definition immediately leads to the differential equations
for the geodesic lines. On the other hand, the geodesic line can be defined as
a curve with the shortest distance between two infinitesimally close points
on the surface. Under this definition the variational statement implies the
requirement for the integral determining the length of the curve of the
surface to have a stationary value.
In statics while looking for the equilibrium form of a chain line (a heavy
homogeneous chain with fixed ends) we can start with the differential equa-
tions of the line obtained from analysis of forces acting on an infinitesimal
element of the chain. Another definition is based on the idea that the centre
670 12. Variational principles in mechanics

of gravity of the sought-for curve must reach its lowest position at equilib-
rium. This implies a variational formulation of the problem in question.
Let

(12.1.1)

be functions of time which are the generalised coordinates describing the


real motion of the material system. Let the system be subject to holonomic
ideal constraints and the active (prescribed) forces be potential forces. The
set offunctions (1) is said to determine the true path of the system whereas
any of the oon configurations

(12.1.2)

admitted by the constraints and infinitesimally close the true path defines
the varied path. In eq. (2) variations 8qs imply arbitrary, infinitesimal, dif-
ferentiable functions of time.
Let L denote the kinetic potential which is the difference between the
kinetic and potential energy. Along the true path

(12.1.3)

is a given functions of time. For a varied path the variation of the kinetic
potential is

(12.1.4)

Here we used the rule "8d = d8" explained in Sec. 1.7.


From all possible varied paths we select those which coincide with the
true path at two fixed (however arbitrarily taken) time instants to and tl,
so that

(12.1.5)

Let us now introduce quantity S referred to as Hamilton's action over


time interval (to, td and defined as follows

J
t,

S = Ldt. (12.1.6)
to

Hamilton's action is the functional determined by the set of n functions


of time qs (t). On the true path S takes a particular value S* and functions
qs (t) and qs (t) are the generalised coordinates and velocities under the real
motion.
Let us proceed to calculate the increment in S when we take one of
the oon varied paths instead of the true one. The value of the kinetic
12.2 The Hamilton-Ostrogradsky principle 671

potential along this varied path should be inserted into eq. (6). Restricting
our consideration to first order values in 8q8 and 8q8 we should replace L
by L + 8L. Then, the increment in Hamilton's action calculated with the
mentioned accuracy is given by the equality
h tl tl

8*+88=j(L+8L)dt= j Ldt+ j8Ldt


to to to

and is called the variation in Hamilton's action. According to eq. (4) we


obtain

(12.1. 7)

Integrating by parts and taking into account the additional condition (5)
imposed on the varied paths we have

(12.1.8)
Hence, variation in action is determined by the following expression

(12.1.9)

where £8 (L) denotes the Eulerian operator (7.1.8).

12.2 The Hamilton-Ostrogradsky principle


Turning to expression (1.9) for variation in action we have two possible lines
ofreasoning. First, from the position of Newton's axioms and D'Alembert's
principle we can say that q8 (t) satisfy Lagrange's equations

£8(L)=0 (s=l, ... ,n) (12.2.1)


under the real motion and thus

88= o. (12.2.2)
Secondly, another point of view is possible, namely, equality (2) is an
independent requirement. This leads to the statement of the Hamilton-
Ostrogradsky principle which reflects a certain property of real motions and
the difference from all the possible motions admitted by the constraints.
672 12. Variational principles in mechanics

Given the functional of the sort (1.6)

J
Xl

J= F(ydx), ... ,yn(X),y~(X), ... ,y~(x),x)dx,


Xo

the set of functions Ys (x) is said to render this functional stationary if the
variation of this functional caused by variations 8ys of function Ys (x) and
calculated up to first order terms in these variations vanishes.
Hence, the Hamilton-Ostrogradsky principle states that Hamilton's action
S has a stationary value for the true path as compared with all arbitrary
neighbouring paths coinciding with the true path at the initial to and final
it time instants.
If we prove that Lagrange's equations (1) are a consequence of equality
(2) then we can state that the Hamilton-Ostrogradsky principle contains
the basic laws of dynamics. The proof is as follows. As variations 8qs are
independent of each other we can adopt the following neighbouring path

8ql = 0, ... ,8qk-1 = 0, 8qk+l = 0, ... ,8qn = 0,


where 8qk =I 0. Equation (2) takes the form

(12.2.3)

Assume now that

(12.2.4)

at t = t*. Then we can find such a time interval (t*o, t*l) that it contains
t* and £k (L) retains its sign. But 8qk is an arbitrary function of time.
Let us choose it so that it retains its sign within the above interval and is
identically equal to zero at to < t*o and tl < t.l. Then,

J J
tl t.l

£dL) 8qk dt = £dL) 8qk dt .


to t.o

The integrand retains its sign and thus the integral does not vanish. This
contradiction means that the assumed inequality (4) does not hold. Hence,

£dL) =0.
This reasoning is valid for any k = 1, ... ,n which completes the proof.
Hamilton formulated the principle of stationary action for the free system
of particles and the system of particles subject to stationary constraints in
12.2 The Hamilton-Ostrogradsky principle 673

his papers on dynamics and optics published in the mid-thirties of the nine-
teenth century. In 1848 this restriction was removed by Ostrogradsky who
not being familiar with the Hamilton papers published in the little known
Transactions of the Irish Academy of Science derived the principle in the
paper on differential equations of isoperimetric problem and generalised
the principle to non-stationary constraints. The fundamental classical pa-
pers on variational principle are collected in [74]. It contains a letter by
Ostrogradsky to Brashman in which the principle of stationary action is
explained in nearly modern terms.
Let us also demonstrate derivation of the canonical equations of motion
in terms of the Hamilton-Ostrogradsky principle. We begin with equality
(10.2.9)
8L
n n
L = L a:-8qs - H = LPsqs - H. (12.2.5)
s=1 qs s=1
Then

(12.2.6)

Equalities
. 8H
qs =- (s= 1, ... ,n) (12.2.7)
8ps
are the first set of the canonical equations and have no relation to mechanics
since they are a form of relationship between the generalised coordinates
and momenta. Thus

(12.2.8)

Taking into account eq. (1.5) we have

J JP JP
tl tl tl

Ps (8qsr dt = Ps8qsl~~ - s8qsdt = - s8qsdt


to to to

and therefore

(12.2.9)

By repeating the derivation which yields Lagrange's equations from eq. (2),
we arrive at the second set of the Hamiltonian canonical equations
8H
Ps=-- (s=l, ... ,n), (12.2.10)
8qs
674 12. Variational principles in mechanics

which completes the proof.


An analogous approach is applicable to the derivation of the Euler-
Lagrange equations (8.1.8). Quasi-velocities Ws and variations of quasi-
coordinates 87f s are introduced into consideration by formulae (1.5.1) and
(1.6.12). Then using expression (4.1.19) for the kinetic energy in terms of
the quasi-velocities and the definition of "the derivative of a function with
respect to quasi-coordinate" (1.5.17) we arrive at the following equality for
the kinetic potential

r
Replacing here 8w s in terms of (87f s byeq. (1.8.3) and taking into account
the rule" 8d = d8" and property (1.9.2) of the three-index symbols, we have
n n

8ws = (87f s r + L L '"Y~twt87fr (12.2.11)


t=l r=l

and moreover

Integrating by parts

aT
-a 87f s
It! - Jh 87f d aT
-a
dt Ws dt
S -
Ws to
to to

and noticing that variations of the quasi-coordinates reduce to zero

(12.2.12)

due to conditions (1.5) and (1.6.12) we obtain

Since quantities 87f s are independent of each other we come to equations of


motion (8.1.8). Relationships (8.1.9) expressing the generalised velocities in
terms of the quasi-velocities should be added to these equations. This ap-
proach allows derivation of equations of motion for the free body by means
of the Hamilton-Ostrogradsky principle. The derivation of the equations of
12.2 The Hamilton-Ostrogradsky principle 675

motion of a rigid body based on relationships between variations of pro-


jections of the angular velocity and time-derivatives of the vector of the
infinitesimal rotation was given in the classical lectures by Kirchhoff [48].
In the case of non-potential forces the relationship analogous to the
Hamilton-Ostrogradsky principle is cast in the form

J J
tl n tl

08 = o'Wdt = 08 + L Qsoqsdt = o. (12.2.14)


to s=1 to

As follows from the above

J L J£8 (L)oq8 dt = L J(;~ -!;~) oq8 dt ,


tl n h n tl

08 = oLdt =-
to 8=1 to 8=1 to
(12.2.15)

and equality (14) reduces to the form

(12.2.16)

which yields Lagrange's equations. If we introduce variations of the quasi-


coordinates and represent the expression for the elementary work in the
form of eq. (5.1.6)
n
O'W = LP807r8' (12.2.17)
8=1

then making use of eq. (13) for variation of action we arrive at the Euler-
Lagrange differential equations of motion.
Attention should be paid to the principal difference of equality (14) from
the previous relationships (2) and (9). The latter require a functional 8
and a search for the necessary conditions of stationarity of this functional,
that is, the mechanical problem was reduced to the problem of calculus of
variation. In contrast to this, eq. (14) contains only the statement that the
value

J J
tl tl

0' R = 08 + o'wdt = (OT - oIl + o'W) dt (12.2.18)


to to

turns to zero, however no functional R exists since there is no value whose


variation equals 0' R. Hence, eq. (14) implies no variational statement of
the problem. Relationship (13) is not a variational notation either, since
its right hand side contains variations of quasi-coordinates.
676 12. Variational principles in mechanics

We have seen that the Hamilton-Ostrogradsky principle is valid for the


systems subject to holonomic constraints under potential and non-potential
forces. It will be. shown in the sequel that this principle is also applicable
to systems subject to non-holonomic constraints.

12.3 On the character of extremum of Hamilton's


action
It is known that the complete differential df of function f (Xl> . .. ,xn ) is
equal to zero at points where this function has a stationary value. The
character of the stationarity (minimum, maximum, absence of minimum
or maximum) is stated by the sign of the second differential d2 f and, in
some exceptional cases, by means of the sign of higher-order terms. Thus,
the condition 88 = 0 is only the necessary condition for stationarity of
Hamilton's action on the true path. The character of the extremum can be
clarified by means of the sign of the second variation 82 8, the minimum
occurring under the condition 82 8 > O. Limiting our consideration to sta-
tionary constraints it is easy to show the presence of a minimum of the
action 8 for a sufficiently small time interval tl - to. To this end, we con-
struct the expression for the increment l:lL in the kinetic potential under
transition from the true path to the varied path

Here the calculation is limited by terms of second order in 8qs and 8cjs.
Under stationary constraints the kinetic energy is a quadratic form of the
generalised velocities and, thus, the last group of terms can be cast as
follows

where T (8q) denotes the result of replacing the generalised velocities qs in


the equation for T by their variations 8qs. Then
12.3 On the character of extremum of Hamilton's action 677

Taking into account that 8q8 (to) °


= we have

J
tl

18q8 (t)1 = 8q8 dt < (38 (t - to),


to

where (38 denotes the maximum ofthe absolute value of 8q8 for to < t < h.
Hence, for a sufficiently small time interval tl - to the term T (8q) prevails
and determines the sign of eq. (1)

82 L ';::j T(8q) > 0,


since the quadratic form T is positive definite. Thus

J J
tl it

82 8 = 82 Ldt ';::j T(8q)dt > 0, (12.3.2)


to to

which completes the proof.


Let us recall that the true path in the Hamilton-Ostrogradsky principle
is defined as the motion of the system between two a priori prescribed
positions Ao and Al given by the generalised coordinates q~ at t = to
and q; at t = tt, respectively. Along the neighbouring paths the system
passes through the same positions at the prescribed time instants. Thus,
there is no reason to think that a neighbouring path exists or it is unique,
if it exists. Indeed, in the theory of differential equations the uniqueness
and existence of the solution under the given initial conditions are proved
(the so-called Cauchy problem). The conditions of Cauchy's problem are
fulfilled in mechanics.
The search for the true path is a boundary-value problem: it is necessary
to determine such initial momenta (or initial generalised velocities) that
the system reaches the terminal position from the initial one in the given
time interval. The problem may have no solution, one solution, several or
even an infinite number of solutions.
An important case is that in which there exist infinitesimally close true
paths between the positions Ao and Al of the system, the paths being
followed at the same time. These two positions of the system are termed
the conjugate kinetic foci. A well-known example is the motion of a free
particle on the surface of a sphere. This motion occurs with a constant
velocity along the great circle of the sphere which is a geodesic line. The
conjugate kinetic foci are the positions of the particle on the ends of the
same spherical diameter since they can be connected by infinitesimally close
great semi-circles, the time needed for covering any semi-circle being the
same under the same initial velocity.
Let us return to inequality (2) which is held for a sufficiently small time
interval h - to. Let ti designate the time instant at which the second vari-
ation 82 8, calculated along the neighbouring path AoH AI, turns to zero
678 12. Variational principles in mechanics

II F

~
Ao

FIGURE 12.1.

for the first time, see Fig. 12.1. The neighbouring path AoH Al is assumed
to intersect the true path AoBAI at instants t = to and t = ti. The action
along the true path AoBAI is denoted by S* . Along any neighbouring path
intersecting the true one at points Ao and Al the action differs from S*
by a second-order value. According to the formulated problem, the action
along the neighbouring path AoH Al is equal to S* with accuracy up to
second-order values since 82 S = 0 along it. Let us show that this neigh-
bouring path is the true path. Let us suppose the opposite, that is suppose
that path AoH Al is not the true one. Then, connecting two sufficiently
close points E and F by the true path ERF we find that the action along
path ERF is smaller than that along EHF. Hence, the action along path
AoERF Al is smaller than along path AoEH F Al and in turn smaller than
S*. But this contradicts the condition that Al is the first position on the
true path AoBAI where the second variation 82 S becomes zero along the
neighbouring paths intersecting the true path.
Therefore, positions Ao and Al are connected by two infinitesimally close
true paths, i.e. they are the conjugate kinetic foci. Along with this, we
proved that the action is a minimum provided that the system reaches
the final position before the kinetic focus of the initial point. This proof
reproduces Jacobi's idea [44] for a particular case of the motion of a particle
on the sphere and was suggested by Whittaker in [95] .
Let Ao and Al be the conjugate kinetic foci. We consider the true path
ABAI Q whose final position Q is reached after the focus Al has already
been passed, see Fig. 12.2. The action along this path is no longer the min-
imum. It follows from the possibility of constructing a neighbouring path,
the action along which is smaller than that along the true path ABAIQ. In
order to prove this , let us take position N on the true path AH Al which
is so close to Q that the action along the true path NTQ is a minimum,
then

Thus

S* SNTQ = SABAlQ = SABA l + SAlQ = SAHA l + SAlQ


SAHN + SNA l + SAlQ > SAHN + SNTQ = SAHNTQ ,
12.3 On the character of extremum of Hamilton's action 679

r-
A

FIGURE 12.2.

that is the action along the true path turns out to be larger than along the
constructed neighbouring path.
This geometric construction allows us to establish the presence of the
minimum of Hamilton's action along the true path which does not pass
through the kinetic focus and the absence of minimum if the true path
passes through the kinetic focus . However this construction does not pro-
vide us with a means for searching the conjugate focus and does not solve
the problem of its existence.
Let us proceed to some detailed analysis, [11]. Expressing L in terms of
the canonical variables
n
L = LPsqs - H (q1, ... ,Qn,P1 , ··· ,Pn)
s=l

(12.3.3)

where the second derivatives of the Hamiltonian function H are calculated


on the true path. For the quadratic form in expression (3) for {)2 L we adopt
the notation

(12.3.4)

which enables equality (3) to be written in the form


680 12. Variational principles in mechanics

Then, taking into account the condition for the neighbouring paths

(12.3.6)

we arrive at the following expression

2
o s = 21 Jhto ~
n [ (.
oPs oqs -
oD ) (.
oops - oqs ops
oD )]
+ ooqs dt. (12.3.7)

In what follows we return to this expression. Now we consider the equa-


tions of motion which are infinitesimally close to the motion under consid-
eration, i.e. the variational equations for the canonical system of equations

. oH .
Ps=--
oH (s=l, ... ,n). (12.3.8)
qS=8' oqs
Ps
Constructing these equations in the way explained in Sec. 11.10 we obtain

(s=l, ... ,n), (12.3.9)

where Xs and Us denote variations of coordinates and momenta, and D is


the above quadratic form (4). Also we notice that for Lagrange's equations

.
Ps=-
oL (s=l, ... ,n), (12.3.10)
oqs

the variational system of equations will be

of . of
Us = oX s ' Us =- (s=l, ... ,n). (12.3.11)
oXs
Here F denotes the following quadratic form

(12.3.12)

As seen from formula (1) the second variation is represented by the same
form F of variables Oqk and Oqk, so that

J
tl

02 S = F (Oql, ... , oqn; Oql, ... , oqn) dt. (12.3.13)


to
12.3 On the character of extremum of Hamilton's action 681

The solution of Cauchy's problem for the system of differential equations


of motion (8) is

qs = qs (t - to, a!, . . . ,an,/31,··· ,/3n ),


Ps =Ps(t-to,a1, ... ,an,/31,··· ,/3n)
(8 = 1, ... , n),} (12.3.14)

ak and /3 k denoting the generalised coordinates and momenta at t = to.


Then, according to Sec. 11.11 the following functions of time

(8,m=I, ... ,n)


(12.3.15)
present the system of particular solutions of linear variational equations
(9).
In as much as equations (14) can be considered as the formulae for the
canonical transformation of variables (ak,/3k) to (Pk,qk) we can use eq.
(10.8.19)

~~ ~~ TJt TJf
D ( q1,··· , qn,P1,··· ,Pn ) = ~~ ~~ TJ~ TJ~ =1
a!, ... , an,/31,··· , /3 n d G 'l9 11 'l9 n
1

(~ (~ 'l9 n1 'l9 n
n
and thus solutions (15) are linearly independent. According to the definition
of the integral of the system of equations we have

~~ (to) = 8sm, TJ~ (to) = 0, (~ (to) = 0, 'l9~ (to) = 8sm. (12.3.16)


We also need equality (10.7.8) for the Lagrange brackets

Having solution (15) and using its property (16) we can cast the integral
of Cauchy's problem for the variational system of differential equations

Xs = j; n
[xm (to) ~~ + Um (to) TJ~], _
(8 - 1, ... , n) .
}
(12.3.17)
Us = L [Xm (to) (~ + Um (to) 'l9~l
m=l

Let us return now to expression (7) for the second variation of action. It
is equal to zero if the equalities
o. 80, o. 80,
(8=1, ... ,n). (12.3.18)
qs = 80ps' Ps = - 80qs
682 12. Variational principles in mechanics

hold. But these equations coincide with the variational system of equations
(9) provided that Xs and Us are replaced by 8ps and 8qs, respectively. The
conclusion obtained above from geometrical reasoning follows here from
the fact that the neighbouring paths along which 82 S = 0 belong to the true
paths which are infinitesimally close to the true path (14).
Let us proceed to the question of the existence of the neighbouring paths
along which 82 S = O. It is clear that a positive answer can be given if
variational equations (18) have non-trivial solutions. By satisfying the first
set of conditions (6) under the above replacement we obtain, due to eq.
(17), that
n n
8qs (t) = L U m (to) ",,;, (t) , 8ps = L U m (to) 79';' (t) (s = 1, ... , n) .
m=l m=l
(12.3.19)

For the forthcoming analysis we need the determinant

8q1 8q1
",t "'1 8131 8f3n
A(t,to) = (12.3.20)
8qn 8qn
"';, "'~ 8131 8f3n
where A (to, to) = 0 as follows from equalities (16).
Turning to the second set of the above conditions we arrive at the system
of linear homogeneous equations
n
8qs (t1) = L u m (to) ",,;, (t1) = 0 (s = 1, ... , n) , (12.3.21)
m=l

which has non-trivial solutions for variations of momenta Um (to) provided


that

(12.3.22)

Here t1 denotes the root of equation A (t, to) = 0 that is nearest to to.
Under condition (22) the unknown parameters U m (to) are found up to
a constant factor. This defines a bundle of paths (19) which are true and
infinitesimally close to the true path under consideration. Hamilton's action
calculated up to second-order terms is the same for all these paths. They
intersect at instants to and it at conjugate kinetic foci whose positions are
given by the formulae

as, qs (t1) = qs (t - to, a1,··· , an, 13 1"" , f3n)


(s = 1, ... , n) . (12.3.23)
12.3 On the character of extremum of Hamilton's action 683

It remains to prove that for any t from interval (to, iJ) when the inequality

~ (t, to) i- 0 (to < t < tt) , (12.3.24)

holds, Hamilton's action has minimum along the true path. In other words,
it is necessary to prove whether the second variation 02 8 is positive for a
finite time interval (24) rather than for a sufficiently small interval as stated
above. It will be proved that the sign of quadratic form F, which is the
integrand in eq. (13), coincides with the sign of the following form
1 n n
"2 L L c sk 8qs8qk = T (8q) > 0, (12.3.25)
s=1k=1
i.e. it is positive definite. The proof for case n = 2 is given in [34] and [33].
We follow the proof suggested in [45] for arbitrary n however the present
proof is essentially simplified.
The consideration is based on substituting into F the new variables
n
8qs=vs+L'Ysr8qr (s=I, ... ,n) (12.3.26)
r=1
and adding the quadratic form
1 d n
"2 dt L Ask8 qs8qk, (12.3.27)
s=1
to the integrand in eq. (13). Because the latter quadratic form is a com-
plete time-derivative and conditions (6) are satisfied, this addition does not
change the value of integral (13). By virtue of eq. (12) equality (13) takes
the form

n n n
CskVsVk + Csk L L 'Y sr'Ykh 8 qr 8Qh + 2CskVs L 'Ykr 8qr +
r=1 h=1 r=1

).sk 8qs8qk + 2Ask8qk (vs + ~ 'Y srOqr) 1 dt (12.3.28)

and our goal is to choose such functions of time 'Y sr and Ask that this
expression becomes

(12.3.29)
684 12. Variational principles in mechanics

where, by virtue of eq. (25), the equality takes place only for Vs = 0 for all
8 = 1, ... ,n. Hence, it is necessary to convince ourselves that all 8qs are
zero, that is 82 S = 0 is possible only along the true path (14). Hence, it is
necessary to satisfy the conditions
n
bsk +L csr'Yrk + Ask = 0,
r=l
n n n
(12.3.30)
ask +2L (brk + Ark) 'Yrs + L L Crh'Yrs'Yhk + >'sk = 0
r=l r=lh=l
(8, k = 1, ... ,n) ,
where the second set of conditions is transformed by means of the first set
to the form
n n
ask = ->'sk + L L Crh'Yrs'Yhk (8, k = 1, ... ,n). (12.3.31)
r=lh=l

The matrix of coefficients Ask of the quadratic form (27) is symmetric


however bsk =I- Aks. Thus, the equations
n
bsk + LCkr'Yrs+Aks =0 (k,8= 1, ... ,n) (12.3.32)
r=l

should be added to the equations. It remains to show, first, that all equa-
tions obtained have solutions and, second, that all 8qs are zero at Vs = o.
Notice that, due to eq. (26),
n
8qs = L 'Ysr 8qr (8 = 1, ... ,n) (12.3.33)
r=l

at Vs = o.
Let us return to some properties of the solutions (15) of the variational
equations. Let us assume that
n n

19:;' = - L lskTJk (8,m=1, ... ,n), (12.3.34)


k=l

where 9sk and lsk are functions of time which can be found from the above
relationship for any fixed 8 under condition (24). Matrix lsk is symmetric
which can be proved by means of eq. (10.7.8)

n n
L L TJkTJ':' (lrk - lkr) = O.
k=lr=l
12.3 On the character of extremum of Hamilton's action 685

Applying condition (24) twice, we obtain

lkr=lkr (k,r=l, ... ,n). (12.3.35)

Let us now relate functions gks and lks with the coefficients of the dif-
ferential equations (11). Substituting the particular solutions 79'; and 'TI':'
into the first set of equations (11) for Us and Xs we have

for any m = 1, ... , n. We can rewrite the latter result as

Returning to eq. (24) we find


n
lsk + bsk + L Csrgrk = 0,
r=l (12.3.36)
n
lks + bks + L Ckrgrs = 0
r=l
From relationships (34) and the second set of equations (11) we obtain

Equating the coefficients of 'TIl: and using eq. (36) we arrive at the equalities
n n
isk + ask - L L Crhgrsghk = 0 (8, k = 1, ... , n) . (12.3.37)
r=lh=l
Comparing eqs. (36), (37) with eqs. (30), (32) and (31) we find that the
latter can be satisfied by taking

(12.3.38)

Therefore, by means of functions gsk and lsk determined from eq. (34)
we carry out the linear transformation of variables (26) by reducing the
expression for the second variation 82 S to the form (29). Equations (33)
which are now cast in the form
n
8qs = Lgsk8qk (8 = 1, ... ,n) (12.3.39)
k=l
686 12. Variational principles in mechanics

possess only a trivial solution under conditions (6). Indeed, from the first
set of equalities (34) it follows that the solution which becomes zero at
t = to is
n

DQs = L CmTJr:', (12.3.40)


m=l

whereas, under the condition of non-vanishing determinant (20), the second


set of equations in (6) yields

Cm=O (m=1, ... ,n). (12.3.41 )

Let us summarise the above.


1. If determinant (20) does become zero in time interval [to, tl], i.e. the
true path does not pass through the kinetic focus, then Hamilton's action
has a minimum and the solution is unique.
2. If .6. (tl, to) = 0, then positions Ao and Al at time instants to and
t l , respectively, are the kinetic foci and there exist other infinitesimally
close paths for which the condition of stationarity of the action holds. The
boundary-value problem has an infinite number of solutions.
In the calculus of variations, conditions (24) and (29) are termed Jacobi's
condition and Legendre's condition, respectively. In mechanics the latter
condition is the requirement of positive definiteness of the kinetic energy
and thus is always met. Jacobi's condition holds along the true paths which
do not pass through the kinetic focus corresponding to the initial position.
We can determine the kinetic foci in the following way. Without loss of
generality we can take to = 0 and qo = O. The motion of the system is
described by the equalities

Qs=Qs(t,p~, ... ,p~) (8=1, ... ,n), (12.3.42)

where p~ denote the initial momenta. In order to find the true path we
should determine them from the system of equations

(12.3.43)

Q; denoting the generalised coordinates at position AI. The Jacobian


Oql Oql
op~ op~
.6. (t) = (12.3.44)
oqn oqn
op~ op~

is identically equal to zero at t = 0 since equations (42) are identically


fulfilled for arbitrary momenta p~ at this time instant.
12.3 On the character of extremum of Hamilton's action 687

When the initial momenta are P~ +8p~, with 8p~ being arbitrary infinites-
imal values, we have

~ 0 ~ 0) 8qs ~ 0 ~ 0
qs ( t I , PI
0
+ uPI' ...
0
'Pn + uPn = qs + 8 0 uPI + ... + 88qs0 uPn + ...
PI Pn
(12.3.45)

where dots denote terms of second order of smallness in 8p~ and higher. If
Jacobian (44) at t = tl (different of t = to = 0) equals zero again, then the
system of equations

8q; ~ 0 8q; ~ 0
8 0 UPI + ... + 8 0 UP n = 0 (8 = 1, ... ,n) (12.3.46)
PI Pn
will have the non-zero solutions 8p~. Then, due to eqs. (45) and (42) we
obtain

qs (tI'P~ + 8p~, ... ,p~ + 8p~) = qs (tI,P~, ... ,p~) +... (8 = 1, ... ,n)
(12.3.47)

Thus, if at t = tl Jacobian (44) is equal to zero, then there are two


systems of the initial values of momenta

p~ and p~ + 8pt
which result in two infinitesimally close paths, namely path (42) and the
path

qs (t,p~ + 8p~, ... ,p~ + 8p~) (8 = 1, ... ,n),


both bringing the system to position Al or to a position which differs from
Al by small values of second order or higher. In this case Al is the kinetic
focus conjugate with the initial one.
Let us prove this for the above mentioned example of motion of the free
particle on a spherical surface. We take the radius of the sphere as unity
and choose the coordinate system in such a way that point A is initially
ca
on equator o = 1f /2, >"0 = 0). Let us denote the projections ofthe initial
velocity on the directions of the tangent to the meridian and equator by
-ao and >'0. Then the particle moves along the great circle inclined to the
equatorial plane by angle i, where

. >'0 -a
cosz = -, sini = --,
w w
w denoting the constant angular velocity. The arc along the great circle is
wt and, as follows from the spherical triangle AMQ, we have

cos V.0 = - --ao SIn


. wt, ·.0
sIn \
v cos /\ = cos wt, •
sIn .0 • \
v sIn /\ = ->'0.sIn wt ,
w w
68812. Variational principles in mechanics

z
N

!I

FIGURE 12.3.

see Fig. 12.3. Therefore


.
sm{)-.-
a{) (.2 ·2
w13 AO sin wt + {) owt cos wt , )
a{)o
. a{) 19 0>"0 .
sm{)-.- - 3 - (- sm wt + wt cos wt) ,
8Ao w
1 aA 19 0>"0
- - ( -tanwt+ - wt
- -)
cos 2 A 019 0 w3 cos 2 wt '
1 aA
3
1
( ·2
{)o tan wt + --2-
.2)
Aowt
.
cos 2 A 8>"0 W cos wt

Jacobian (44) is equal to

~ (t) = w; sinwt
w
and becomes zero at t = 7r /w and A = 7r, which correspond respectively to
the position {)1 = 7r /2 and point F diametrically opposite to point A.
As a second example let us consider the problem of free vibration of the
oscillator. The motion is given by the equality

q = 0: cos wt + ~ sinwt, (12.3.48)


w
where 0: and (3 denote the initial values of the generalised coordinate and
velocity, respectively, and w is the natural frequency. Due to eq. (44) we
12.3 On the character of extremum of Hamilton's action 689

have

~()= oq = sin wt (12.3.49)


t 0(3 w'

that is the kinetic focus is reached at time instant t1 = 7f / W at position


q1 = -Q. Indeed, the expression

q = Qcoswt + ..!.. ((3 + 8(3) sinwt, (12.3.50)


w

with 8(3 denoting an infinitesimal value, is the solution of the differential


equation for the oscillator and represents the true path which is infinites-
imally close to the true path (48) and intersects it at the kinetic focus.
Applying the theorem of variational calculus we can say that eq. (56) de-
scribes a bundle of extremals emanating from point qo = Q and intersecting
each other at the kinetic focus.
This example allows us to carry out the calculation confirming the above
geometric proof of the absence of a minimum of Hamilton's action along the
path possessing the kinetic focus corresponding to the initial position. Let
the final point of the true path (48) be passed at instant t = 7f / W + T. Let
us construct the neighbouring path from two parts: the first part coincides
7f
with the true path for 0 ~ t ~ - - T whilst the end of this part is connected
w
by the true path with the final point of the prescribed true path (48), so
that this part is passed in time interval 2T. Hamilton's action along the
second part is a minimum if 2T < 7f / w. Then denoting the generalised
coordinate along the neighbouring path by q* (t) we have

O<t<~-T ,
q* (t) = { qi (t) , 7f
- -
w7f (12.3.51)
q2 (t) , - - T ~ t ~ - +T,
W W

qi (t) being given byeq. (50) within the above time interval. The positions
of the final points of path qi (t) and the true path (48) are given by the
equalities

q~ (~ - T) = -QCOSWT + ~ ((3 + 8(3) sinwT = q2 (~ - T), (12.3.52)

q (~+ T) = -QCOSWT - ~(3SinWT = q2 (~+ T) . (12.3.53)

FUnction q2 (t) is the solution of the differential equation of vibration

q2 = C 1 coswt + C 2 sinwt,
690 12. Variational principles in mechanics

where constants 0 1 and O2 should be determined by means of conditions


(52) and (53). We arrive at the system of equations

0 1 COSWT - O2 sinwT aCOSWT - ~ (;3 + 8;3) sinwT,


W
1 (.l •
0 1 COSWT + O2 sinwT aCOSWT + -fJsmWT,
W

from which it follows that

where 2WT < 1r. Hence,

q2 (t) (a - ~~ tanWT) COSWT + ~ (;3 + ~;3) sinwt,


(~ - T~ t ~ ~ + T) . (12.3.54)

It remains to calculate Hamilton's actions along the true and neighbouring


paths
7r

J(
-+7"
W
1
8 = 2 q2 -W 2 q2) dt,
o
7r
--7" -+7"

J J
W W

8* = 8i + 82 = ~ (qf - w2qf) dt + ~ (q2 2 - W 2q22) dt.


o 7r
--7"
W
(12.3.55)

To simplify this calculation we notice that the action during the time in-
terval (to, h) for the harmonic vibration

q = Acoswt + ~Bsinwt
W

is equal to

J L
tl

8 =~ (q2 - w2q2) dt = sinw (t1 - to) X

to
[(B2 - w2A2) cosw (to + t1) - 2ABwsinw (to + h)] . (12.3.56)
12.3 On the character of extremum of Hamilton's action 691

By using this formula, it is easy to calculate expression (55), to have

sinwt 2 2 2 . (8,8)2
S = - - [(,8 - a W ) COSWT - 2a,8wsmwT] , S* = S - --tanwT.
2w 4w
Hence,

1 2
S - S* = 4w (8,8) tanwT > 0, (12.3.57)

which completes the proof.


Construction of determinant (20) assumes knowledge of the solution of
Cauchy's problem of the system of differential equations of motion. When
the general solution of this system having 2n constants is known

(12.3.58)

(which is not the solution of the initial-value problem) then the search of
the kinetic foci is as follows. Equations (58) determines a certain true path
Co for fixed Cl, ... , C2n. A set of true paths C' infinitesimally close to Co
corresponds to values Ck + 8Ck of these constants. Along these paths

8
L
2n
8qs = 8 qs 8c r (s = 1, ... , n).
r=l Cr

If paths C' intersects Co at instants t = to and t = h then the correspond-


ing configurations of the system q~Ol and qFl are the conjugate kinetic foci.
We come to the homogeneous system of 2n linear equations

L
2n (88~; ) 8cr = 0, (s=l, ... ,n),
r=l 1

having non-trivial solution for 8cr if the system's determinant vanishes, i.e.

~ (t, to) = =0. (12.3.59)

The time instant tl is found from this equation, see [11].


692 12. Variational principles in mechanics

12.4 Application to non-holonomic systems


If the system is subject to non-holonomic constraints prescribed by the
equations
n
L askqk + CY s = 0 (s = 1, ... ,l), (12.4.1)
k=l
then the variations describing the neighbouring path should satisfy, in ad-
dition to eq. (1.5), the following equations
n
Lask8qk =0 (s = 1, ... ,l). (12.4.2)
k=l
These equations determine an instantaneously fixed configuration of the
system admitted by the constraints. If we consider not an instantaneously
fixed configuration but the motion of the system, then the constraint equa-
tions (1) should hold also for varied generalised coordinates along the neigh-
bouring path. In other words, along with eq. (1) the relationships
n

L ask (ql + 8ql,··. ,qn + 8qn, t) [qk + 8qk] +


k=l
as (ql + 8ql,· .. ,qn + 8qn; t) = 0 (s = 1, ... ,l) (12.4.3)

must hold. Taking into account eq. (1) and restricting the analysis to the
first-order terms we find

(12.4.4)

On the other hand, differentiating equations (2) we have

(12.4.5)

Subtracting eq. (5) from (4) and using the rule" d8 = 8d" we arrive at the
equations

~ L~ (aa
L a
S k _ aa sr )
a . 8qr + L
qk ~ (aa s sr
at ) 8qr
a _ aa = 0, }
k=l r=l qr qk r=l qr
(s = 1, ... ,l),
(12.4.6)

which are referred to as the conditions for the kinematic feasibility of ad-
jacent motion, see [77]. From a class of variations 8qk subject to conditions
12.4 Application to non-holonomic systems 693

(2) imposed on adjacent configurations these equations select a narrower


class of variations admitting motions adjacent to the true one. Under the
integrability conditions

(r, k = 1, ... , n; s = 1, ... , l) (12.4.7)

equations (6) are fulfilled automatically. Hence, any adjacent configura-


tion is feasible under an adjacent motion provided that the constraints are
holonomic.
Let us turn to the relationship

J
tl n

LEk (L) 8qkdt = 0 (12.4.8)


to k=l

expressing the Hamilton-Ostrogradsky principle. There is no need to re-


quire that variations 8qk, defining transition to the neighbouring path and
satisfying conditions (2) should also satisfy conditions (6). The statement
of the fundamental equations of dynamics (see Secs. 6.3 and 1.6) does not
compare the true motion with the motion admitted by constraints. It com-
pares only two positions of the system, namely the position in the real
motion with an infinitesimally close position admitted by constraints at
the same time instant. The same interpretation of the concept of variation
should also be adopted for the Hamilton-Ostrogradsky principle if we view
it as a consequence of the basic statements of mechanics.
The derivation of the equations of motion for the non-holonomic systems
from relationship (8) is now beyond question. We can pursue two ways. The
first way is to express variations 8ql, ... , 8ql in terms of the remaining in-
dependent variables 8ql+l, ... , 8qn, substitute them into eq. (8) and obtain
n - l equations by equating the factors of the independent variations to
zero. Then we obtain the equations of motion (7.10.9). The second way
is to apply the method of Lagrange's multipliers which implies that each
equation in (2) is multiplied by an undetermined multiplier As and the sum
of these products (equal to zero) is placed into integral (8). Then we arrive
at the equality

(12.4.9)

Then we choose the multipliers AI, ... , Al in such a way that l expressions
in the square brackets turn to zero, the remaining n - l variations in the
integrand are independent and the factors in front of them must be zero
too. Finally we obtain the equations of motion in the form of (7.1.6)
I
EdL)+LAsask=O (k=1, ... ,n), (12.4.10)
s=1
694 12. Variational principles in mechanics

which should be considered together with the constraint equations (1).


On introducing quasi-velocities Ws and casting the equations for the non-
holonomic constraints in the form
n
Ws = L askqk = 0 (s = 1, ... , l) (12.4.11)
k=l
we can set conditions (2) imposed on the variations of quasi-coordinates as
follows
n
87rs=Lask8qk=O (s=1, ... ,l). (12.4.12)
k=l
In eq. (2.13) the terms vanish if their subscripts sand t take values
1, ... , l. No condition is imposed on variations 81'1 s for s = l + 1, ... , nand,
hence, the coefficients of these variations can be equated to zero. This leads
to the equations of motion (8.1.13)

~ ,:aT
}
+ ~ ~ rt aT Wt
L L's'::}
'Y _ aT = ,arr
,: } :} (s=l+1, ... ,n).
dt uWs Z
t=
uW
+1 r=l r u7r s u7r s

(12.4.13)

The following kinematic relationships


n
qs= L bsk w k (s=1, ... ,n) (12.4.14)
k=/+l
should be added to the above equations. It is significant that the constraints
should not be taken into account when constructing the expression for
the kinetic energy since equations (13) contains terms with ~T for r =
UW r
1, ... , l also. It is also an error if we take into account the presence of the
constraint and remove l generalised velocities (ll, ... , qZ from L before the
derivatives needed for construction of equations of motion (10) are taken.
This questions has been discussed in Sec. 8.7.
Integral (8) containing variations 8qk related by eq. (2) is not variation
of an integral. The same is true for integral (9). Similar to the system with
non-potential forces the requirement for the integral to be equal to zero
expresses the Hamilton-Ostrogradsky principle rather than a variational
principle. The derivation of the equations of motion of the system subject
to the non-holonomic constraints based on this idea is given by Holder, see
[39].
The question arises as to whether the Hamilton-Ostrogradsky principle
in the case of non-holonomic constraints can be treated as a constrained
problem of the calculus of variations, namely an isoperimetric Lagrange's
12.4 Application to non-holonomic systems 695

problem

J
tl

S = Ldt (12.4.15)
to

subject to constraint (1), [34].


In order to solve the problem we introduce the function

(12.4.16)

with /-Ls being the undetermined multipliers, and construct equations for
the extremals for the functional

J
h

R= <Pdt. (12.4.17)
to

The variation of this integral is given by

(12.4.18)

The second integral on the right hand side vanishes by virtue of the con-
straint equations (1). Integrating by parts and taking into account the
condition for choice of the neighbouring paths (1.5) we have

The right hand side of eq. (18) is cast as follows


696 12. Variational principles in mechanics

After a little algebra


n
L (qk 8ask - ask 8qk) + 8as
k=l

= t [t (~aSk
r=l k=l qr
- ~asr)
qk
qk 8qr + (~as - f)~;r) 8qr ]
qr U
(12.4.20)

This expression becomes zero if we take into account the conditions of


the kinematic feasibility of the adjacent motion (6). Denoting its as As we
obtain

J J JL
tl tl h I n

8R 8if>dt = 8Ldt - As L a sk8qk dt

-t, I [Ek (L) t,a. A.]6 d ~


to to to s=l k=l

+ k Qk t o.

We arrived at equality (9) which implies that we have obtained the equa-
tions of motion (10). Hence, these equations are obtained from the con-
strained problem of the variational calculus provided that the conditions
of the kinematic feasibility of the adjacent motion are taken into account
along the sought-for extremals. Generally speaking, this attempt to keep
the variational formulation of the Hamilton-Ostrogradsky principle does
not lead to the desired result since the conditions of the kinematic feasi-
bility of the adjacent motion can be consistent with conditions (2) and the
constraint equations (1) only if these equations are integrable. This can be
shown for a simple example of the motion of a material point under the
non-holonomic constraint
(12.4.21 )
where the functions of the coordinates aI, a2, a3 are considered as projec-
tions of vector a. This example is studied in [39] and [85]. The velocity of
the particle remains perpendicular to the lines of the vector field a. The
constraint is integrable (holonomic) if the integrating multiplies M (x, y, z)
exists. Then the expression
M (a1dx + a2dy + a3dz)
is the complete differential and hence
rotMa = Mrota - gradM x a = O.
The integrability condition obtained by removing M is

a· rot a = a1 (~~ - ~~ ) + a2 (~:1 - ~~ ) + a3 (~~ - ~~ ) = o.


(12.4.22)
12.4 Application to non-holonomic systems 697

Equation (6) can be written as follows

(a + Da) . (v + DV) - a· v - a· (Drr - it· Dr = V· Da - it· Dr = O.


(12.4.23)

We have
oa oa oa
Da = -Dx + -Dy + -Dz = 'Va· Dr, it ='Va . v,
ox oy oz
where 'Va = grad a denotes the following tensor of second rank, see Sec. 9.8

oal oa2 oa3 oa2 oal oa3 oal


2 oa1 -+- -+-
ox ox oy ox OZ
/J~ /Jffz /J!f3 1
-
oal
-+-
oa2
2 oa2 oa3 oa2
-+- +
oy oy oy 2 oy ox oy oy OZ
oal oa2 oa3 oal oa3 oa2 oa3
-+- -+- 2 oa3
OZ OZ oz oz ox oz oy oz
oa2 oal oa3 oal
0 --- ---
ox oy ox OZ
1 oal oa2 oa3 oa2
- --- 0 --- = defa+ e,
2 oy ox oy OZ
oal oa3 oa2 oa3
--- --- 0
oz ox oz oy
Here def a and e denote the symmetric and skew-symmetric parts of tensor
'Va. Equation (23) is nOw cast in the form

v· Da - it· Dr = (v· def a· Dr - Dr· def a· v) + (v· e· Dr-Dr· e· v) = o.


The first parenthesis vanish since for any symmetric tensor

a· Q . b = b . Q . a.
The expression in the second parenthesis can be written as follows

v . e . Dr-Dr· e .v (v x ~ rot a) . Dr- (Dr x ~ rot a) .v


v . (rot a x Dr) ,

thus, the conditions of the kinematic feasibility of motion (23) is expressed


by the equality

v· (rot a x Dr) = o. (12.4.24)

In the example under consideration equation (2) takes the form

a· Dr = 0, (12.4.25)
698 12. Variational principles in mechanics

which enables us to represent Dr as the vector product

Dr = a x b,
where b denotes a vector whose direction is arbitrary. Then, using the
constraint equation (21) we can write condition (24) in the form
y. (rota x Dr) y . [rot a x (a x b) 1
y . ab . rot a - y . ba . rot a = -y . ba . rot a =0.
If we choose vector b in such a way that it is not perpendicular to y, then

a· rota = 0, (12.4.26)
that is, the simultaneous fulfillment of the two conditions (25) and (24)
leads to the requirement of integrability of constraints.

12.5 Equations of motion of distributed systems


We refer to the material systems whose configuration cannot be described
by a finite number of generalised coordinates as distributed systems. If
we restrict our consideration to one-dimensional systems, examples are
a flexible inextensible rope, a string and an elastic rod. The Hamilton-
Ostrogradsky principle provides us with a mostly simple and natural way
of constructing the equations of motion for such systems. They are partial
differential equations with two independent variables, namely time t and
the coordinate prescribing the position of the particle (cross-section) along
the body (rope, string, rod).

12.5.1 Vibration of a hanging chain with a mass on the end


Let us consider a flexible, inextensible, homogeneous, heavy chain of length
I, see Fig. 12.4. End 0 is fixed and a particle of mass m is attached at end
N. The position of a generic point M is given by the abscissa a = OM in the
equilibrium position. We consider planar vibration of the chain and denote
the projections of point M on axes Ox and Oy by u and v, respectively.
Functions u (a, t) and v (a, t) determine both motion of any point of the
chain (for a fixed a) and the form G' of the chain (for a fixed t) . Projections
of the displacement of point N, i.e. mass m, are equal to u (I, t) and v (I, t).
The condition of inextensibility of the chain results in a certain relation-
ship between functions u and v which is a constraint equation. We obtain
it by expressing the condition that the length da of the chain element re-
mains unaltered. Position M' of point M at time instant t is given by its
Cartesian coordinates

x = a + u (a, t), Y = v (a, t). (12.5.1)


12.5 Equations of motion of distributed systems 699

o y

!
X

FIGURE 12.4.

This is the parametric equation of curve G' for a fixed t. The differential
of arc dO" is da, so that

This yields

(1 + au)2
aa + (av)2
aa = I , au
aa = _~
2
[(au)2
aa + (av)2].
aa (12.5.2)

The analysis is limited to small deflections of the chain from the vertical,
i.e. we adopt that u, v and their derivatives with respect to a and tare
small. Let a prime and a dot denote derivatives with respect to a and t,
respectively. Then if v, v' , v are considered as quantities of the first order
of smallness, then, by virtue of eq. (2) u, u', U are second-order quantities.
Since (u,)2 is of the fourth order of smallness we can neglect it in eq. (2)
and write eq. (2) as follows

(12.5.3)
700 12. Variational principles in mechanics

The kinetic energy of the system to the above accuracy is

J(u2+v2)da+~m[U2(I,t)+v2(l,t)]
I

T = ~p
o

J.2v da+ 2mv


I

21 P 1.2 (l,t,
) (12.5.4)
o
where P denotes the mass of a unit length of the chain. The potential energy
of the weight is given by

II = -glpxc - mgxN,
where Xc and XN stand for the coordinates of the centre of mass of the
chain and mass m, respectively. They are as follows

J+ J +J J
I I I I

lxc = (a u) dO" = ada uda = I; + uda, XN = 1+ u (I, t).


o 0 0 0

Let us remove u from these expressions by means of the constraint equation.


Taking into account that u (0, t) = 0 we obtain

J J
I I

u (I, t) = u' (a, t) da = -~ V,2 da,

o 0

J = aul~- J = -~l J da+~ J J


I I I I I

uda au'da V,2 av Pda = -~ (I - a) V,2 da.


o 0 0 0 0

Omitting the immaterial constant we find

II = 21 gp J I

(I - a) v ,2 da + 2mg
1 J ,2
I

v da = 21 gp J+
I

(I h - a) v ,2 da,
0 0 0
(12.5.5)
where
m
-=iI=fJ l
p
and fJ denotes the ratio of the end mass to the chain mass.

1
The kinetic potential L takes the form

L ~ ~P { [0' - 9 (I + l, - a) v"l da + I,;,' (I, t) } . (12.5.6)


12.5 Equations of motion of distributed systems 701

Now the variational problem derived from the Hamilton-Ostrogradsky prin-


ciple is stated as follows: among all continuous functions v(a, t) having con-
tinuous derivatives with respect to a and t for 0 :::; a :::; land t > 0 and
satisfying the boundary condition

v (0, t) = 0, (12.5.7)

it is necessary to find a function which makes Hamilton's action

S = J
tl

to
Ldt = ~P J {IJ[1;2 -
t,

to
dt
0
9 (l +h - a) V,2] da + hi} (l, t)
}
(12.5.8)

stationary. Here the "true path" is determined by the sought-for function


v (a, t) whereas the neighbouring path by v+bv (a, t). The variation bv (a, t)
means an arbitrary continuous function in the domain of definition 0 :::; a :::;
land t > 0 which satisfies the conditions

bv (a, to) = 0, bv (a, td = 0, bv (0, t) = 0 (12.5.9)

and has continuous derivatives bv' and bv. The first and the second equal-
ities in eq. (9) express condition (1.5) of the choice of the neighbouring
paths in the Hamilton-Ostrogradsky principle whilst the last one is the
consequence of the boundary condition (7) since point 0 is immovable.
Variation of the action is given by

bS = J {IJ
P t, dt
to 0
[vbv - 9 (l + it - a) v'bv'] da + lIV (l, t) bv (l, t) }.
Next, by means of integration by parts it is possible to remove derivatives
bv' and bv. To this end, the rule" db = bd" is applied. Using eq. (9) we
obtain

J+ J
I I

(l h - a) v'bv'da (l + h - a)v'bvl~ - bv :a (l +h - a)v'da


o o

J
I

hv'(l,t)bv(l,t) - bV:a (l+h -a)v'da


o
as well as

J J J
tl tl tl

vbvdt = vbvl~~ - vbvdt = - vbvdt.


to to to
702 12. Variational principles in mechanics

The Hamilton-Ostrogradsky principle yields

DS p J
to
dt{; [-v+g:a (l+h -a)v'] 6vda-
0

h [gv' (l, t) + v (l, t)] Dv (l, tn. (12.5.10)

Since variation Dv considered as function of time is arbitrary, the integrand


of the integral over t must be zero. However Dv (a, t) as well as Dv (l, t) can
be arbitrarily prescribed. Hence, each expression in the square brackets
must be equal to zero. This leads firstly to the partial differential equation
of motion for the chain

v = 9 :a [(l + h - a) v'] (12.5.11)

and, secondly, to the boundary condition at the end x = l

h [gv' (l, t) + v (l, t)] = O. (12.5.12)

The boundary condition at the end a = 0 is given by eq. (7). Let us


notice that in the case of nO mass at the end, i.e. h = 0, condition (12)
is replaced by the requirement that v and its derivatives with respect to a
and t are bounded.
As in any mechanical problem, the initial conditions determining the
position and velocity at the initial instant of time must be prescribed.
They have the form

v(a,O)=f(a), v(a,O)=g(a), (12.5.13)

where f (a) and 9 (a) are continuous functions of a, such that f (0) = 0 and
9 (0) = O.
Briefly, the route to solving the boundary-value problem obtained is as
follows. As the theory of differential equations in mathematical physics
suggests function v (a, t) is sought as a sum of particular solutions Vk (a, t),
each of them being the harmonic vibration referred to as the principal
vibration

(12.5.14)

Frequency Wk is not yet determined. Function <I>k (a) termed the normal
mode determines the amplitude distribution along the chain and is the
solution of the following differential equation
2
[(l +h - a) <I>Ua)]' + Wk <I>da) = 0, (k = 1,2, ... ) (12.5.15)
9
12.5 Equations of motion of distributed systems 703

subject to boundary conditions.

<I>k (0) = 0, (k = 1,2, ... ) . (12.5.16)

Introducing the new independent variable (J by means of the equality

1 9 2
Z+h-a=--(J
4w%
differential equation (15) is reduced to Bessel's equation

Its solution can be cast in the form

where J o ((J) and No ((J) are Bessel functions of the first and second kind,
respectively. Noticing that

we can write the boundary conditions (16) in the form

GlkJO (Xk) + G2k No (Xk) = 0, }

Glk [J6 (xk) + ~Xi:Jo (Xi:)] + G2k [N6 (xi:) + ~Xi:No (Xi:)] = 0,
(12.5.17)

where

(12.5.18)

The system of equations (17) has a nontrivial solution for Glk, G2k if its
determinant is equal to zero. Then we arrive at the transcendental equation

Jo (Xk) [N6 (xi:) + ~xi:No (Xi:)] - No (Xk) [J~ (xi:) + ~xi:Jo (Xi:)] = 0.
(12.5.19)

which has an infinite number of roots. All of the roots are simple and
determine the required natural frequencies by means of eq. (18).
704 12. Variational principles in mechanics

The solution Vk (a, t) can be written down in the form

(12.5.20)

so that

<I>k (a) = No (Xk)JO ( Xk V 1 - l : lJ - Jo (Xk) No (XkV1 - l : lJ .


The constants Ak and Bk are determined by means of the initial conditions
(13)

(12.5.21)

These formulae are actually expansions of the prescribed functions f (a)


and 9 (a) in terms of the normal modes <I>k (a).
The solution is based on the property of the generalised orthogonality of
the system of functions <I>k (a) expressed as follows

J
I

<I>k (a) <I>m (a) da + h <I>k (l) <I>m (l) = 0 (k i- m). (12.5.22)
o
Indeed, turning to differential equation (15) we can write

:% J<I>da) <I>m (a) da + J[(l + h - a) <I>Ua)] , <I>m (a) da


I I

= O.
o 0

After double integration by parts and using eq. (15) together with the
boundary condition (16) we obtain

~ (w;;, - wi) [[ <Pda) <Pm (a) da + f, <Pdf) <Pm (I)1~ 0,


which yields relationship (22) for m i- k. For m = k we introduce the
notation

J<I>~
I

Nm = (a) da + h <I>~ (l). (12.5.23)


o
12.5 Equations of motion of distributed systems 705

Then, taking into account that due to eqs. (22) and (23)

Jf
I

(a) cJ>m (a) da + f (l) cJ>m (l) h


o

~ A, [f '!>,<Pm da + I, <Pk (l) <Pm (l)] ~ A".Nm ,


we obtain

(12.5.24)

and by analogy

Bm = - -
xkNm
2
f¥[ Jg
l +h
--
g
l

o
(a) cJ>m (a) da + hg (l) cJ>m (l)1 (12.5.25)

The solution is determined completely. Substantiation of the validity of the


operations performed and the proof of convergence of the series obtained
are beyond the scope of the present book.
The numerical work can be carried out for any particular ratio JL, however
it is considerably simplified for the case in which the end mass is absent,
i.e. JL = 0 or h = O. Then

1 2
l- a = -ga
4 '
and a = 0 at the lower end of the chain. The solution of Bessel's equation
is

since, if we kept Bessel function of the second kind, the solution would be
unbounded at a = O. The frequency equation obtained from the boundary
condition (7) is

(12.5.26)

The first three roots are given by

Xl = 2.4048, X2 = 5.520, X3 = 8.654,


706 12. Variational principles in mechanics

so that

Wl = 1. 2024 1f, W2 = 2. 760 1f, W3 = 4.327 If·


The normal modes take the form

iPda)=Jo(XkV1-y) (k=1,2, ... ). (12.5.27)

Equalities (22) and (23) yield

j Jo (XkR) Jo (xmR) da= {


o
(k -I m)
(k = m) ,

where
Nl = 0.269l, N2 = 0.116l, N3 = 0.074l.
The solution is thus the following series

v(a, t) = ~Jo (XkV 1 - y) (AkCOS x; 1ft + Bksin X; 1ft) ,


(12.5.28)
where

JJ(a)Jo (XkV1 - y) da,


I

Ak = ~k
o (12.5.29)

Xk~k f£ J (a) Jo (XkV 1 - y) da.


I

Bk = 9
o
The motion of the chain is a superposition of the principal vibrations,
each of them being a harmonic oscillation of frequency Wk. The form of
the k - th principal vibration is given by function iP k (a). According to
expression (27), this function has, besides a = 0, another k -1 roots, given
by the equalities

a~) = l (1 - X~)
x
(m = 1,2, ... ,k - 1). (12.5.30)
k

These determine the nodal points of the k-th vibration mode. For example,
the second mode has a single node

a (2)
1
_
- l ( xi) ~
1 - x~ ~ 0.81l,

the third mode has two nodes


ai3) = 0.923l, a~3) = 0.595l
and so on. The first mode has no nodes at all.
12.5 Equations of motion of distributed systems 707

FIGURE 12.5.

12.5.2 Vibration of a rotating elastic rod


This problem has been considered in Sec. 9.10. The end 0 of the elastic
homogeneous rod with constant cross-section is clamped to a disc of radius
R rotating about the immovable axis Oz with constant angular velocity w.
The other end of the rod x = I is free, see Fig. 12.5. We are required to
construct the differential equation of bending vibrations in planes Oxy and
Oxz , [81].
Denoting the vector of displacement of point M in the rod as u and its
projections on the axes of the rotating coordinate system Oxyz as u, v, w
we obtain the following expression for the radius vector r' of point M'

r' = i l (u + R + a) + hv + hw.
In Fig. 12.5 point M' shows the actual position of point M of the unde-
formed rod. In the latter equation a = OM and u, v, ware the sought-for
functions of a and t. The condition of inextensibility of the rod is derived
by analogy with that for a chain and has the form

U
, = --1(,2 ,2)
v +w . (12.5.31 )
2

The absolute velocity of the point M is equal to

where Vr denotes its relative velocity. Then we obtain

vx=u-wv, Vy=v+w(u+R+a), vz=w ,


708 12. Variational principles in mechanics

and the kinetic energy of the rod is given by

J J
I I

T ~P (V;+v~+v~)da=~p da[w 2v 2 +v 2 +2wv(R+a)+


o 0

2w 2 (R+a)u+w2 (R+a)2 +W2].

Here only terms of second order of smallness are kept (please, notice that
u is a second order quantity whereas v and ware first order quantities).
Since

J J
I I

(R+a)uda (R+ ~a) aul~ - (R+ ~a) au'da


o o

-~ J[( + ~l)
I

R l - ( R + ~a ) a] (V,2 + W,2) da
o
we can cancel out the nonessential constant term in the expression for T,
to obtain

J
I

T ~p da{w 2v2 +v 2 +2wv(R+a)-


o
w2 [R (l - a) + ~ (l2 - a2)] (V,2 + W,2) + w2} . (12.5.32)

The potential energy of bending is known to be equal to

J(
I

II =~ Elzv l/2 + Elywl/2) da, (12.5.33)


o
and Hamilton's action takes the form

JJ
tl I

S ~p dt da [v 2 + w2 + w2v 2 + 2wv (R + a) - (12.5.34)


to 0

w2 (l - a) 2R + l + a ( v ,2 + w,2) - - v ,2 - -
Elz w ,2] .
Ely
2 p p
The required functions v (a, t) and w (a, t) are subject to the boundary
conditions

v (0, t) = 0, w (0, t) = 0, v' (0, t) = 0, w' (0, t) = 0. (12.5.35)


12.5 Equations of motion of distributed systems 709

are obtained from the condition 88 = °


The differential equations of vibration and the boundary conditions at x = l

tl I

88 = P J dt J da ['118'11 + w8w +w 2v8v +w (R + a) 8'11-


to 0

w2 (l - a ) 2R + l + a (v '8v '
2
+ w '8')
w - Elz
-
P
v"8 v II - -
Ely
w"8 w"]
P
=.°
(12.5.36)

Integrating by parts and taking into account the equalities

8v (a, to) = 0, 8w (a, to) = 0, 8v (a, tl) = 0, 8w (a, tl) = 0,


8v' (a, to) = 0, 8w' (a, to) = 0, 8v' (a, h) = 0, 8w' (a, h) = 0,

we can recast eq. (36) in the form

h I
88=pJdtJda { _v+w 2v+w 2 [(l_a)2R+ l+a v'] ' }
_E:zv IV 8v
2
to 0

.. 2R+l+a 'EI}
+ P Jh dt JI da {
-w + [w2 (l- a) 2 w'] - --;-wIV 8w
to 0
tl

+ Elz J dt [v"' (l, t) 8v (l, t) - v" (l, t) 8v' (l, t)]


to
h
+ Ely J dt [w"' (l, t) 8w (l, t) - w" (l, t) 8w' (l, t)] = 0.
to

Finally, we arrive at the following partial differential equations and bound-


ary conditions: for vibration in plane Oxy

..
V +-
Elz
v IV - w2 v - w2 [(l - a ) 2R + l + a v ,] = 1o}
P 2'
°
(12.5.37)
v (0, t) = 0, v' (0, t) = 0, v" (l, t) = 0, VIII (l, t) =

and for vibration in plane Oxz

.. Ely IV -w 2 [(l -a ) 2R + l + a W ,] I =
w+-w
P 2 '
° } (12.5.38)
w (0, t) = 0, w' (0, t) = 0, w" (l, t) = 0, Will (l, t) = 0.
Integration of these equations was carried out in Sec. 9.10.
710 12. Variational principles in mechanics

, I
IJ..o._-~- 20 • I
I
I V

!fo
o

FIGURE 12.6.

12.5.3 Vibration of a chain line


We consider small oscillations of a heavy homogeneous inextensible chain
about its equilibrium position. The ends of the chain are fixed at the same
level, see Fig. 12.6.
The equilibrium form of the chain is generally called the chain line. This
form is obtained if we consider the variational problem for the minimum
of the potential energy of the weight of the chain

J J
a a

II= ,), yda=,), yVl+y,2dx.


-a - a

Here ,)" ±a and da = VI + y,2dx denote the weight per unit length of
the chain, abscissas of the fixed point, and the arc element of the sought-
for curve y (x), respectively. The latter equation can be understood as the
formula for Hamilton's action S if we replace x by t. Then we deal with
"motion" of the single degree-of-freedom system with the kinetic potential
L = YVI +y,2 ,
which does not contain the independent variable x. The "equation of mo-
tion" , which is Euler's equation for the variational problem under consid-
eration, has the energy integral (7.2.5). In our case it is given by

y' fJL _ L =
fJy' VI y+ y,2 = -Yo,

where the constant Yo denotes the value of y at the lowest point of the
equilibrium form (where y' = 0). Hence,

y2 = Y5 (1 + y") ,
12.5 Equations of motion of distributed systems 711

and differentiating yields

y" - .JL
Y5 -- 0
.
The solution, which is even and equal to Yo at x = 0, is
x
y = yocosh-. (12.5.39)
Yo
Then we have
x
y' = sinh~, da = cosh -dx, (12.5.40)
Yo Yo
and the length of the chain is

J _l_ = sinh (2a ._l_) .


a

I = cosh ~dx = 2yo sinh .!!:.. or (12.5.41 )


Yo Yo 2yo I 2yo
-a

The unknown value _l_ should be found from this transcendental equation
.
£or a gIven parameter T.
2yo
2a
In what follows we will need the formulae
dx 1 . a = tanh -x ,
cos a = -d = x ' sIn (12.5.42)
a cosh- Yo
Yo
where a denotes the angle of the tangent to axis Ox. From them we obtain
dx x da 1 1
- =yocosh-, x' (12.5.43)
da Yo da P yocosh-
Yo
with p designating the radius of curvature of the chain line.
Let us proceed to construct the differential equation of vibration of the
chain. We denote the vector of the displacement of the chain point from the
equilibrium position by u, whilst /-l,v,j3 are its projections on the tangent,
main normal and binormal to the chain line, respectively. Then

u = /-IT + vn + j3h, r = ro + u, (12.5.44)

where ro and r are the position vectors of the chain point in the initial
and actual position, respectively. Taking the derivative with respect to the
curvilinear coordinate a and using Frenet's formulae for the planar curve
we obtain
712 12. Variational principles in mechanics

For the inextensible chain Idrl = du, which implies that

( 1- ~ + 8J-l)2 + (~+ 8v)2 + (8(3)2 = l.


P 8u p 8u 8u

After simple manipulations we have

~= 8J-l +~ [(~ + 8v)2 + (~ _ 8J-l)2 + (8(3)2]. (12.5.45)


P 8u 2 p 8u p 8u 8u

This condition for inextensibility of the chain is actually the constraint


equation for the problem under consideration. Neglecting terms of second
order and higher, we write this constraint equation in the form
v
(12.5.46)
p
Keeping the second order terms we can cast the constraint equation as
follows

(12.5.47)

We proceed to construct the equation for the potential energy. By virtue


of eq. (44) the vertical coordinate of the chain point is

ry = Y + J-l sin a + v cos a,


where y denotes the coordinate of this point on the chain line at the equi-
librium. Then, taking into account the inextensibility of the chain we have

IT "I 1 rydu = "I 1 ydu + "I 1 (J-l sin a + v cos a) du

1
000

ITo + "I (J-lsina + vcosa) pda,


-000

with ITo being the value of the potential energy in the equilibrium position.
The integration limits correspond to the values of a at the fixed ends.
Replacing v by eq. (47) yields

1
000

IT - ITo "I (J-lSina + ~~ cos a ) pda +


-000

1
"2"1 lao cos a [( J-l + 8
8 2; )2 + (8(3)2]
8a da.
-000
12.5 Equations of motion of distributed systems 713

The linear term must vanish as the potential energy has a stationary
value in the equilibrium position. Indeed, replacing sin a, cos a, p by their
expressions (42), (43) we obtain

7
-<>0
(MSina + =~ cos a ) pda
-<>0
/
<>0 (
Mtanh -
x
Yo
aM) cosh -dx
+ 7lYo
uX
x
Yo

because M= ° at x = ±a. Thus,

(12.5.48)

which confirms that the potential energy has a minimum in the equilibrium
position. The kinetic energy of the chain is given by

T =

(12.5.49)

While deriving this result we used relationship (46) and neglected the
fourth-order terms.
Now, entering the non-dimensional variables

= rit , ~=~
VYo
T
Yo
and denoting the derivatives with respect to these variables by a dot and
a prime, respectively, we obtain the following expression for the kinetic
potential
~o

L = ~')' / [Ji,2 cosh~ + Ji,p cosh3 ~ - (CO~h~ + (12.5.50)


-~o

~o

Mil cosh~ + M' sinh~)2] ~ + ~')' / (il cosh~ - ,8,2) d~.


-~o

Here eqs. (42) and (43) are used and ~o denotes the value of ~ at x = a.
The required functions M and ,8 satisfy the boundary conditions
at ~ = ±~o M = 0, M' = 0, f3 = 0, (12.5.51)
714 12. Variational principles in mechanics

the second condition being a consequence of equality (46).


Equating variation of the action to zero, i.e. 88 = 0, and integrating by
parts in order to remove the derivatives of variations 8J.L and 8(3 by means
of the boundary conditions for the variables, we obtain the differential
equations for small vibration of the chain

p, - p," cosh2 .; - 3p,' cosh'; sinh'; + 2J.L' sinh'; + 5J.L" cosh'; +


4P,''' sinh'; + J.LIV cosh'; = 0, (12.5.52)

,Bcosh'; - (3" = 0, (12.5.53)

Taking

J.L = M (.;) sin (At + a) , (3 = B (.;) sin (wt + a) ,


we reduce the problem of determining the frequencies and the form of
the vibration about the equilibrium to two boundary-value problems: for
vibration in the plane of the chain

M1V cosh'; + 4M'" sinh'; + (5 + A2 cosh';) M" cosh';+ }


(2 + 3A 2 cosh';) M' sinh'; - A2 M = 0, (12.5.54)
M (±';o) = 0, M' (±';o) = 0,

and for vibration perpendicular to the plane of the chain

B" + w 2 B cosh'; = 0, B (±';o) = 0. (12.5.55)

Paper [80] is devoted to determining the frequencies of free vibration of the


chain in its plane. Differential equation (54) is constructed there for the
independent variable a.

12.6 Approximate determination of natural


frequencies and normal forms
The previous section was concerned with constructing the differential equa-
tions for small vibration of one-dimensional bodies about the equilibrium
position in the potential field. The equations obtained are partial differ-
ential equations for unknown function v (x, t) , with x and t denoting re-
spectively the coordinate of the point in the equilibrium configuration and
time. The solution had two steps. First, we introduced a family of partial
solutions

(12.6.1)
12.6 Approximate determination of natural frequencies and normal forms 715

each being the harmonic oscillation referred to as the principal oscillation.


The natural frequencies Wk and normal modes Vk (x) must be determined.
On substituting eq. (1) into the differential equation, the variables x and t
are separated and determination of Vk (x) is reduced to the homogeneous
boundary-value problem. The solution of this problem under trivial bound-
ary conditions provides us with such values of w~ which ensure non-trivial
functions Vk (x). The matter of the second step was to fulfill the initial
conditions. The solution of the problem was represented by series in terms
of the normal modes Vk (x), constants Ck and Dk were determined by the
prescribed initial deflection and velocity.
The principal realisation of these operations can be proved, however the
representation of the solution in terms of well-known functions is possible
only for a limited range of problems. For instance, it wa.."l not possible in
subsections 12.5.2 and 12.5.3 since the differential equations for the normal
modes are not known. This gives rise to the problem of using approxi-
mations to determine the frequencies and modes of vibration. In practical
applications it is important to know a few first frequencies and modes, pre-
dominantly, the fundamental one. In the framework of this approach the
distributed system is replaced by a system with a finite degrees of free-
dom which is equal to the number of required normal modes. The initial
conditions need to be specified at the same number of points.
Let us start with consideration of the system having a finite number of
degrees of freedom and vibrating about the position of stable equilibrium
in the potential field. In this case the kinetic and potential energies are,
respectively, quadratic forms of the generalised velocities and generalised
coordinates with the constant coefficients

1 n n
II = "2 LL
s=lk=l
Cskqsqk· (12.6.2)

It is assumed that, under specially chosen initial conditions, the system can
vibrate in the following way

qs = C s sin (wt + a) (8 = 1, ... ,n), (12.6.3)

where all of the generalised coordinates have the same phase wt + a. Such
vibrations are referred to as principal vibrations. The problem is to ob-
tain the system of algebraic equations for wand C s using the Hamilton-
Ostrogradsky principle. The differential equations of motion are as follows

n
£s (L) = L (askqk + cskqk) = 0 (8 = 1, ... ,n). (12.6.4)
k=l
716 12. Variational principles in mechanics

Substituting expression (3) for the generalised coordinates yields the re-
quired relationships
n
L Gk (-W 2a sk + Csk) = 0 (8 = 1, ... ,n). (12.6.5)
k=l

However, the idea is to avoid constructing the differential equations and


to find the required values directly from the Hamilton-Ostrogradsky prin-
ciple by prescribing the form of the solution [22].
Actually expressions (3) prescribe the true path whilst the harmonic
oscillations with frequencies and amplitudes infinitesimally close to the
required one describe the neighbouring path

q~ (t) G~ sin (w't + a') = qs + bqs


qs + sin (wt + a) bGs + Gsbacos (wt + a) + Gstbwcos (wt + a).
Hence,

bqs = bGs sin (wt + a) + (Gsba + G tbw) cos (wt + a)


8 (8 = 1, ... ,n)
(12.6.6)

The Hamilton-Ostrogradsky principle should be written in the form

J
tl

bLdt = 0, (12.6.7)
to

which is equivalent to the requirement that

J
h

bS = 0, S = Ldt (12.6.8)
to

for fixed to and tl' Provided that to and tl depend on the varied quantities,
the above statements are no longer equivalent. Then, taking into account
that the integration limits are varied, we have

J
h

bS = bLdt + (L)t=tl btl - (L)t=to bto, (12.6.9)


to

and requirement (7) is reduced to the form

(12.6.10)

The same form of the Hamilton-Ostrogradsky principle is suggested in [95].


12.6 Approximate determination of natural frequencies and normal forms 717

Let us assume that t1 and to differ by one period, i.e.


21f
t1=tO+-, (12.6.11)
w
By virtue of the relationship

J 2:. J[s (L) I5qsdt +


t, n t,

I5Ldt = -
to s=l to

~ (:~)[ t=t, I5qs (td - (:~) t=to I5qs (to) 1= 0, (12.6.12)

we can adopt that

I5qs(td-l5qs(to) =0 (s=l, ... ,n), (12.6.13)

due to the periodicity of aa~ rather than consider that each of these vari-
qs
at ions is zero. Then relationship (12) leads to Lagrange's equations which
is required. Due to the periodicity of L and equalities (11), we can cast eq.
(10) in the form

(12.6.14)

Returning to eq. (6) we have

I5qs (t1) = I5Gs sin (wto + a) + Gsl5a cos (wto + a) + Gshl5w cos (wto + a) ,
I5qs (to) = I5Gs sin (wto + a) + Gsl5a cos (wto + a) + Gst ol5w cos (wto + a) .

For condition (13) to be satisfied for to - t1 1= 0 we should adopt that along


the neighbouring paths
1f
wto +a ="2. (12.6.15)

Substituting the generalised coordinates (3) into the expression for L


yields

(12.6.16)

Here rand U are quadratic forms obtained from T and II by replacing qs


and qs by Gs

(12.6.17)
718 12. Variational principles in mechanics

They are referred to as the amplitudes of the kinetic and potential energies.
Then we obtain

JL~t
h

8 = = ~ (w 2 r - U) (12.6.18)
to

and

88 = 7r (r + ~) 8w + ~ (w 2 8r - 8U). (12.6.19)

Noticing that, due to eqs. (16) and (15),


27r 27r
-2 8w
(L)t=t Ow -U-
=
w2 8w,
we can write the Hamilton-Ostrogradsky principle (14) in the form

(12.6.20)

Let us make use of the energy integral

w2 r cos 2 (wto + a) + U sin 2 (wto + a) = const,


where the constant on the right hand side is determined by the left hand
side at t = to. Then we have

or

(12.6.21 )

and it is possible to be represent relationships (20) as follows

(12.6.22)

Thus, if we introduce the quantity

R= -8=w
W 2
r- U, (12.6.23)
7r
its variation 8' R is zero for the fixed w, i.e.

8'R 2
= w 8r - 8U = ~ ( ~ 8r 8U) 8Gs
w2 8G - 8G = o. (12.6.24)
s=1 s s

Since variations 8Gs are arbitrary we obtain the system of equations (5)
by means of eq. (17). The expression for R does not contain the initial
12.6 Approximate determination of natural frequencies and normal forms 719

phase a which is the same for the principal vibration for all coordinates.
The phase a is the arbitrary constant of the particular solution (3).
We arrived at the rigorous system of finite relationships which define
the principal vibration (3) by means of the requirement 0' R = 0, quantity
R being calculated for fixed w. This result is eventually equivalent to the
Hamilton-Ostrogradsky principle, however R is not Hamilton's action and
the variational requirement (24) is not the expression for the Hamilton-
Ostrogradsky principle in the form (2.2). Let us notice that the value of R
along the true path becomes zero by virtue of eq. (21).
Let f * and U* be calculated for the arbitrarily taken values O*s of quan-
tities Os. The value

(12.6.25)

is referred to as the square of the first (fundamental) frequency by Rayleigh.


In passing we notice that squares of the second, third and higher frequencies
can be found provided that certain conditions are imposed on O*s. It can
be proved, however we do not dwell on the proof, that the value of w;
obtained by the latter equation is not smaller than w2 , with w being the
true fundamental frequency. More precisely, w;
is greater than 2 and it w
is equal to w when all the coefficients O*s are equal to Os, i.e. when they
2
define the first normal mode.
Equality (24) is the basis for explaining Ritz's method which is an approx-
imate method for calculating the principal vibrations. Coefficients 01> ... ,
On are not considered to be independent of each other. They are taken to
be linear homogeneous forms of the k parameters ILl' ... , ILk (k < n)
k
Os = LqsrlLr (8 = 1, ... ,n), (12.6.26)
r=l
where the n x k matrix qsr is assumed to be chosen properly. Then f and
U become quadratic forms of parameters ILr and equality (24) takes the
form

Since OlLr are arbitrary we come to the homogeneous system of equations

W2 {Jf _ {JU
{J {J
=0 (r = 1, ... , k) (21.6.28)
ILr ILr
which are linear in ILr.
The requirement of a zero determinant leads to the frequency equation
which is an equation of the k - th power of w2 and has, in general, k
720 12. Variational principles in mechanics

different values wi, ... ,w% which are all positive. For each value of w; there
is a system of values lL~r), ... ,1L~r) determined up to an arbitrary factor.
Then, by virtue of eq. (26) we can determine the coefficients C~r) (up to
an arbitrary factor) corresponding to the normal mode with frequency wr .
In the framework of the described approach the initial system with n
degrees of freedom is replaced by a fictitious system with k degrees of
freedom. Experience has shown that we can obtain frequencies and normal
modes which are very close to the true ones provided that the values of qsr
are properly chosen, that is they are suitable for satisfactory approximation
of the normal modes. For example, for k = 2 one obtains reliable value of wi
and the first normal mode. If one takes k = 4 then one expects satisfactory
accurate information about w~ and the second normal mode as well.
The above is applicable to the systems with distributed parameters. Now
we will consider problems for which the kinetic and potential energies are
the following functionals

I I

T = ~/ p (x) F (il, i/) dx, II = '12 / <P ( v, v I ,v" ) dx, (12.6.29)


o o

where p (x) is the mass per unit length and F and <P are quadratic forms
of their arguments. Assuming

v (x, t) = V (x) sin (wt + a) , (12.6.30)

we obtain expression (16) for L where

I I

f = ~/ p(x)F(V, V')dx, U= ~/ <P (V, Vi, V") dx. (12.6.31 )


o o

The value of R is determined by eq. (23). Integration by parts yields the


following expressions for variations of the amplitudes of the kinetic and
potential energy.

8f ~/ p (x) (~~ 8V + ::,l5V I


) dx (12.6.32)
o

of d
-1 /1 [ p(x)---p(x)- 1 of
OF] 8Vdx+ -p-8V II
2 OV dx oV' 2 OV' 0 '
o
12.6 Approximate determination of natural frequencies and normal forms 721

'2 J
I
1 (8<I> 8V 8<I>, 8<I> ")
8U 8V + 8V,8V + 8V,,8V dx
o

1
'2
J I
(8<I> d 8<I>
8V - dx 8V'
d2
+ dx2 8V"
8<I»
8Vdx +
o
1 [( 8<I>
'2
d 8<I»
8V - dx 8V"
8<I>
8V + 8V,,8V
,] IIo· (12.6.33)

Then we obtain

,
8R =
1
'2
J[
I

w
2 (8F d 8F) 8<I>
P8V - dx P8V' - 8V
d 8<I>
+ dx 8V'
d2 8<I>]
- dx2 8V" 8Vdx
o

+ '2
1 [( 2 8F 8<I>
w P 8V' - 8V'
d 8<I»
+ dx 8V"
8<I>
8V - 8V,,8V
,] II 0 = 0 (12.6.34)

and the problem is reduced to the homogeneous differential equation

2 (8F d 8F) 8<I> d 8<I> d2 8<I> (12.6.35)


W P 8V - dx P 8V' - 8V + dx 8V' - dx 2 8V" = 0

subject to two sets of boundary conditions at x = 0 and x = t. The first


set is: either
2 8F 8<I> d 8<I>
w P8V' - 8V' + dx 8V" = 0
(12.6.36)
or
V=O; (12.6.37)
whereas the second set is: either
8<I>
8V" = 0 (12.6.38)
or
v' =0. (12.6.39)
Conditions (36) and (38) express the force boundary conditions whilst eqs.
(37) and (39) describe the geometrical boundary conditions.
In the framework of the approximate analysis we search, by analogy with
eq. (26), for the normal mode in the form
k
V = LJLr'Pr (x), (12.6.40)
r=l
722 12. Variational principles in mechanics

where 'Pr (x) are some functions satisfying the geometrical boundary con-
ditions. Substituting this expression into formulae (31) and estimating the
integrals we obtain rand U as quadratic form of parameters J.Ll' ... ,J.Lk.
Then we construct equations (28) which yield k frequencies wi, ... ,w% and
k sets of parameters J.L~r), ... ,J.Lr) which in turn allows us to determine the
normal modes Vr (x) up to an arbitrary factor.
Instead of the present method of direct calculation of rand U with
further construction of equations (28) we turn our attention to equality
(34) and replace variation OV as follows
k
oV = L 'Pr (x) OJ.Lr· (12.6.41 )
r=l

Then we obtain equations (28) by equating the coefficients of the indepen-


dent variation OJ.Lr to zero. The result is

Here V should be replaced by its expression (40). This modification of


Ritz's method was suggested by Galerkin.
Let us notice that the non-integral terms in eq. (42)

1 [(
"2
2 () F (}<[>
w P(}V' - (}V'
d (}<[»
+ dx (}V"
(}<[>,]
'Pr (x) - (}V,,'Pr (x)
II
0
(12.6.43)

vanish. Some of these vanish due to the geometrical boundary conditions


since functions 'Pr (x) are subject to these conditions. The remaining terms
vanish provided that 'Pr (x) are subject to force boundary conditions. Then
equation (42) takes the form

1
"2
f[
I

w
2 ( () F d () F ) (}<[> d (}<[> d2 (}<[>]
P(}V-dxP(}V' -(}V+dx(}V'-dx 2 (}V" 'Pr(x)dx=O
o
(r = 1, ... ,k) , (12.6.44)

Contrary to this, the force boundary conditions in Ritz's method (28),


as follows from eq. (42), are automatically included in the requirement of
stationarity of functional R.
12.7 Examples of approximate calculation of natural frequencies and forms 723

The Galerkin equations are easy to construct. Given the differential equa-
tion

\II (V, V', V", ... ) = 0,


the sought-for function V is replaced by its approximation (40), functions
<Pr (x) satisfying all the boundary conditions. Then by means of the follow-
ing relationships

J
I

\II(V,V',V",···)<Pr(x)dx=O (r=I, ... ,k),


o
whose number coincides with the number of the taken functions <Pr (x),
we determine the unknown parameters. Galerkin's method in this form is
applicable to many problems whereas the applicability of Ritz's method
is limited to differential equations obtained by means of the variational
principle. The convergence and accuracy of the solutions obtained by Ritz's
and Galerkin's methods are studied in many papers and books, e.g. [46]
and [67].

12.7 Examples of approximate calculation of


natural frequencies and forms
12.7.1 Vibration of a hanging chain with a mass on the end
The amplitudes of the kinetic and potential energies, eqs. (5.4) and (5.5),
are equal to

J+
I

U = %gp (l it - a) V 2 da, (12.7.1)


o
hence, by virtue of eq. (6.23) we have

(12.7.2)

Let us search V in the form of the polynomial

(12.7.3)

where functions

<Pk(a)=(yf (k=I,2,3)
724 12. Variational principles in mechanics

satisfy the geometrical boundary condition V (0) = O. After elementary, but


bulky, calculation we obtain the following quadratic form of parameters JLk

R = ~ pw 2 {JLi (~ +)() +JL~ (~ +)() +JL~ (~ +)() +


2JLIJL2 (~ +)() +2JL2JL3 (~ +)() +2JL3JLI (~ +)() -
1~2 [JLi (~ +)() +JL~ (~ +~)() +JL~ (130+~)() +
2JLIJL2 (~+){) +2JL2JL3 C+~){) +30 2JL3JLI (~+){)]}, (12.7.4)

where )( denotes the ratio of the end mass and mass of the chain (designated
by JL in Sec. 12.5). Equations (6.28) are easy to construct now. Equating
their determinant to zero yields cubic equations for w 2 , all roots being
positive. Then we obtain three modes Vr (x), among them VI (x) and V2 (x)
being sufficiently accurate. As the calculation is very cumbersome in the
case )( =I=- 0 we restrict our attention to the first approximation by keeping
only the term in eq. (3)

This corresponds to replacing the chain by a rigid rod. Then

and the approximate square of the frequency is given by


2 3 9 1 + 2){
w = -----. (12.7.5)
2 I 1 + 3){
Clearly, factor JLI remains undetermined. If )( = 0 we have

WI = Iff = 1.225/f:,

whereas the exact solution of Sec. 12.5 yields the factor 1.2024. Under the
adopted approximation the result is deemed to be sufficiently accurate.
Taking the deflection of the chain end, both in the exact and approximate
expressions for the vibration mode to be equal to unity, we obtain
a
VI (a) = l'
so that at the midspan

~I (~) = Jo (~) ~ 0.4, VI (~) = 0.5.


12.7 Examples of approximate calculation of natural frequencies and forms 725

The discrepancy in the modes is rather considerable. By means of the sec-


ond approximation for x = 0 we come to the following system of equations

(12.7.6)

The frequency equation

( "31 -
g) (1"5 - 3lw2
2lw2
g) - (1 1 g )2 = 0
"4 - "3 Iw2
has the following roots

WI = If 1.2025, W2 = 1f3.037.

The first approximate frequency nearly coincides with the exact value. How-
ever the second frequency is rather inaccurate as the exact solution yields
2,760.
Inserting the numerical value of WI into eq. (6) we obtain the system of
equations
-0.018JLi1) + 0.0282JL~1) = 0, 0.0282JLi1) - 0.0441JL~1) = 0,

suitable for determination of the coefficients JLi1) and JL~l) of the first mode.
Setting the system determinant to zero yields
JLi1) ~ 1.56JL~1),
and the approximate expression for the first mode is given by

VI (T) = 2.~61.56 ( T+ ~:) ,


which is normalised in such a way that the displacement of the end x = I
is equal to unity.
The table below shows the numerical values of the exact and approximate
expressions for the first mode

a/I ~l (a) VI (a)


0 0 0
0.25 0.178 0.177
0.50 0.398 0.403
0.75 0.670 0.678
1.00 1.000 1.000

Good agreement is observed. In order to obtain the second frequency


and the second mode to satisfactory accuracy it is necessary to keep four
terms in expression (3).
726 12. Variational principles in mechanics

12.1.2 Vibration of a rotating elastic rod


The comparison is restricted to the bending vibrations in the Oxy plane.
Using the expression for oS of Sec. 12.5 we obtain

.' R 1
~ p {(w" + >.') V (a) + w' [(1- a) 2R+21- a v, (a)],
E;z VIV (a)} da + Elz [VIII (l) OV (l) - v" (l) ov' (l)] = 0, (12.7.7)

where ).2 denotes the square of the required eigenfrequency. Taking

a a2
oV' = 2 [2 DILl + 3 T3 oIL2'
(12.7.8)

we meet the geometrical conditions at the clamped end of the rod. Inserting
the latter equation into eq. (6) and equating the coefficients of variations
DILl and 0IL2 to zero we arrive at the equations

(12.7.9)

°
since V IV (a) = under the adopted approximation.
To begin with, we take w 2 = 0, i.e. we consider a non-rotating homo-
geneous cantilevered rod whose frequencies and modes are well-known. It
allows us to estimate the accuracy of the assumed approximation. The
system of equations (9) reduces to the form

(12.7.10)

where
12.7 Examples of approximate calculation of natural frequencies and forms 727

Having constructed the frequency equations we find the smaller root

{Ei:
A = 3.53y 7' (12.7.11)

whilst exact solutions renders the factor to be 3.51560. If w f:. 0 we come


to the following frequency equations

[ ~Z
5
- 4 -1/1 (~+ ~ R)] [~z
15 3 l 7
- 12 -1/1 (~+
35
~ R)]
10 l
-
['61 z - 6 -1/1 (1 +
12
3R)]2
10 T = 0,
_ pl4
1/1 - EIz W
2
, (12.7.12)

which enables estimation of the influence of angular velocity on the funda-


mental frequency of free vibration of the rod.

12.7.3 Oscillation of the mathematical pendulum


In the expression for the potential energy of the pendulum

<p2 <p4 )
II = mgl (1- cos<p) = mgl ( 2 - 24 + ...

1ft t.
we keep the above terms and introduce the non-dimensional independent
variable which is denoted below as Then, the kinetic potential, up
to the constant factor, is as follows

L = ~cp2 _ ~<p2 + <p4. (12.7.13)


2 2 24
The motion of the pendulum under the initial conditions

at t =0 <p = <Po, cp = o. (12.7.14)


is required.
Within the linear approximation the solution has the form

<p = <Po cos t. (12.7.15)


If the last term in eq. (13) is retained, the frequency of vibration changes
and the solution gains a triple frequency harmonic. Under the same initial
conditions (14) we set

<p = (<Po + a) cos At - a cos 3At. (12.7.16)


Using the Hamilton-Ostrogradsky principle we can find A and a. We have

{j<p = {ja (cos At - cos 3At) - t{jA [(<Po + a) sin At - 3a sin 3At] ,
728 12. Variational principles in mechanics

so that 8'P is zero at t = to = 0 and t = h = ~ = T. The upper limit in


the expression for the action depends on the sought-for parameter A, that
is relationship (6.14) should be used
7r
88 - (L)t=T8T = 88 + (L)t=o A28A = o. (12.7.17)

Then we have

if = -A ('Po + a) sin At + 3Aa sin 3At (12.7.18)

and thus

Then, inserting eqs. (16) and (18) into eq. (13), we find

(12.7.19)

It will be shown later that the parameter a has the order of 'P~. With this
in view we have omitted terms higher than second order in a and products
of a 2 and 'Po in the above expression. Furthermore we have

Cancelling out the terms proportional to a 2 , two equations for a and A are
obtained

2+ 2'Poa + A1 (2
'Po -'Po + 2'Poa -
2
5 4)
48 'Po 0,

'Po + lOa - 1 ( 'Po + 2a - 12


A2 1 'Po 3) = O.

Keeping only the lowest order terms in 'Po in the solution we obtain
12.8 Hamilton's principal function 729

and the motion of the pendulum under the adopted approximation is as


follows

4? = 4?o cos ( 1 - ~~) t+ ~; [COS ( 1 - ~~) t- cos 3 (1 - ~~) t] .


(12.7.20)

12.8 Hamilton's principal function


Let us assume that the integral in Hamilton's action has a variable upper
limit and express the kinetic potential in terms of Hamilton's function H
n
L = LPsqs - H (q1, ... , qn,P1,··· ,Pn; t). (12.8.1)
s=1
Then

(12.8.2)

and the variation of this integral is given by

(12.8.3)

By using the canonical equations of motion


8H . 8H
qs = 8ps' Ps=-- (s=l, ... ,n), (12.8.4)
8qs
we transform equality (3) to the form

8S = J
t n t

to
L (Ps8qs
s=1
+ Ps8qs) dt = J
to
n
dLPs8Qs
s=1
=
n
L (Ps8qs -
s=1
f3 s 8as ).

(12.8.5)

Here as and f3 s denote the initial values of the generalised coordinates and
momenta

at t = to qs = as, Ps = f3 s (s = 1, ... , n). (12.8.6)

Next let us recall definition (10.7.5) of the canonical transformation and


expression (10.8.2) for the variation of the generating function of the canon-
ical transformation of the type V (q, Q, t). Then considering S as a function
730 12. Variational principles in mechanics

of the actual and initial values of coordinates, time t and the fixed time
instant to

(12.8.7)

we can conclude, based on equality (5), that this function is the generating
function for the canonical transformation defined by the relationships
as
aqs = Ps, (s = 1, ... ,n). (12.8.8)

After solving these equations for the coordinates qs and momenta Ps


which implies a non-trivial Gessian

G= (12.8.9)

we come to the equalities

(12.8.10)

which provide us with the integral of Cauchy's problem for the system of
the canonical equations of motion. It has been proved in Sec. 10.12 that
relationships (10) present the formulae for the canonical transformation
of the initial values of the generalised coordinates and momenta to their
actual values. In the present book, this celebrated statement is proved by
another way and we see that the generating function of this transformation
is Hamilton's action calculated for the variable upper limit and expressed
in terms of the initial and actual values of the generalised coordinates and
momenta as shown in eq. (7). Hamilton referred to the constructed function
S as the principal function. All of the information about the motion which
can be obtained from the solution of Cauchy's problem, i.e. 2n equations
(10), is contained in the solitary principal function.
Given an explicit expression for the principal function we can immedi-
ately, by means of formulae (8), give the answer to the question as to what
initial momenta should be applied to the system so that the system arrives
at the prescribed position at the prescribed time instant. After Hamilton,
"all mathematical dynamics is reduced to the study and search for the
principal function" .
It follows from equalities (2), (7) and (8) that

dS as n as n n as
dt = at + L 8iJ.s = LpsiJ.s - H = L Ei"";iJ.s - H.
s=1 qs s=1 s=1 q
12.8 Hamilton's principal function 731

The sums on the left and right hand sides cancel out and we obtain

~~ +H(ql, ... ,qn';~' ... ,:~;t) =0. (12.8.11)

Hence, the principal function satisfies the Jacobi equations and represents
the compete integral (10.13.27) for which the constants at, ... , an are equal
to the initial values of the generalised coordinates ql, ... , qn.
Given the solution of Cauchy's problem, the calculation due to formula
(2) yields the expression for the action S in terms of the initial values of
the generalised coordinates, momenta and time. One obtains the principal
function S by removing the initial momenta 13 k from this expression. The
initial momenta should be found from the first set of equations (10).
Another way of constructing the principal function is based On any com-
plete integral of Jacobi's equation

(12.8.12)

where 11, ... "n are the constants. The additive constant In+l is present
since function V appears in Jacobi's equation only in terms of its deriva-
tives. Let us determine this constant such that V = 0 at t = to and qs = as.
Then

V=\II(qt, ... ,qn,'l,··· "n;t)-\II(al, ... ,an,'l,··· ,'n;tO).


(12.8.13)
Let us take the constants in the first system of equalities in (10.13.28) to
be equal to zero, then we arrive at the following system of equations
8\11 (ql, ... , qn, 11,··· , In; t) 8(\IIal' ... ,an,'l,··· "n;tO)

0, (k
8'k
1, ... , n) .
8 'k
= (12.8.14)
The generalised coordinates qt, ... , qn are obtained from this system in
terms of the initial values al, ... , an, time and constants 11, .. · , In. Ex-
cluding the latter from eqs. (13) and (14) we come to the complete integral
of Jacobi's equation which depends on time, and actual and initial values of
the generalised coordinates. This will be the Hamilton principal function.
To illustrate the construction of the principal function we consider mo-
tion of a particle in a vacuum. Using expressions (10.12.9) for the integral
of Cauchy's problem we obtain

J(±2 +
t

S = ~ y2 - 2gy) dt
to

= ~ [(±6 + ya) (t - to) + ~g2 (t - t o)3 - 2gyo (t - to)2 - 2gyo (t - to)] .


732 12. Variational principles in mechanics

Inserting the values of the initial momenta from the first set of equations
(10.12.9)

. x -xo . y-yo 1
x ---
t - to ' Yo = - - + - 9 (t - to) ,
0- t - to 2
into the latter equation yields

s= 2 (t ~ to) [(X - XO)2 + (y - YO)2 - 9 (y + Yo) (t - to)] - ~: (t - to)3 .


(12.8.15)

With the help of formulae (8) we can find the actual and initial momenta.
It is easy to prove that S satisfies Jacobi's equation.
Construction of S by means of the complete integral of Jacobi's equation
is more difficult. This equation, in the problem under consideration, has
the form

oV
-at + -21 [(oV)2 -oy + 2gy1
-ox + (oV)2 = 0

and its complete integral, in the form of eq. (13), is as follows

v = -h(t-to)+(3(x-xo)-
1 [( 2h - (3 2 - 2gy )3/2 - (2h - (3 2 - 2gyo )3/2] . (12.8.16)
3g

If we denote

then the system of equations (14) is reduced to the form

(3
x - Xo = - (T]o - T]) = (3 (t - to).
9

Since T]2 - T]6 = 2g (y - Yo) we find that

y - Yo 1 y - Yo 1
T] = -- - - 9 (t - to) , T]o=--+-g(t-to) (12.8.17)
t - to 2 t - to 2

and moreover
x-xo
(3 --, (12.8.18)
t - to
1 (x - Xo ) 2 1 1 (y - yo) 2 1 2 2
h 2 t - to + 2g (y + Yo) + 2 t _ to + 8"g (t - to)
12.9 Asynchronous variation 733

It remains to substitute these expressions into eq. (16), then we arrive


at the obtained value (15).
In the case of harmonic vibration of the oscillator we have

q = qo cos w (t - to) + Po sin w (t - to) ,


w
and the principal function is as follows

s= 2. ~
smw t - to
) [( q2 + q5) cos w (t - to) - 2qqo] . (12.8.19)

It is easy to prove relationships (8) and (11).


The conjugate kinetic foci can be found with the help of the principal
function by means of the equation
1
C =0,

where C denotes the Gessian (9). The simplest way to derive the latter
equation is as follows. Differentiating the second set of equations (8) with
respect to (3m and taking into account that their left hand sides depend on
(3m only in terms of the generalised coordinates qk we have

(12.8.21)

These relationships can be written in the form of products of the determi-


nants
-1 0 0
0 -1 0
CD. (t, to) = = (-It, (12.8.22)
0 0 -1

where D. (t, to) is determined byeq. (3.20) and reduces to zero in the con-
jugate kinetic foci by virtue of eq. (3.22). Thus we arrive at relationship
(20). For instance, from eq. (15) it follows that the kinetic foci are absent
in the case of motion of the particle in the homogeneous gravitational field.
In the case of the oscillator motion, formulae (19) and (20) yield the time
instant of reaching the kinetic focus which has already been obtained by
means of eq. (3.49), see Sec. 12.3.

12.9 Asynchronous variation


It is not necessary to make comparison of the system configurations along
the true path and one of the adjacent paths at the same time instant, as was
734 12. Variational principles in mechanics

the case in the previous analysis. In other words, prescribing the system
configuration in its true motion by the generalised coordinates qs (t) we can
define an infinitesimally close, admissible by the constraints, neighbouring
motion by function q; (t + D.t). The difference

(12.9.1)

is the synchronous variation introduced above. Taking into account only


values of first order of smallness, we have

q; (t + D.t) = q; (t) + ti; (t) D.t = qs (t) + 8qs + tis (t) D.t, (12.9.2)

so that

q; (t + D.t) - qs (t) = D.qs (t) = 8qs + tis (t) D.t (8 = 1, ... , n). (12.9.3)

This equality defines the asynchronous variation denoted by the symbol D..
Clearly, it can be calculated for any function of time

(12.9.4)

In particular,

(12.9.5)

The value D.t in eq. (3) is an arbitrary differentiable infinitesimal function


of time. Hence,

(D.f)- = (8ft + jD.t + j (D.t)- , (12.9.6)

so that

(12.9.7)

Therefore, the operations D. and d are not commutative, in contrast to 8


and d.
Applying formula (4) to the integral

J
t

Fdt,
o
we arrive at the equality

J J
t t

D. Fdt = 8 Fdt + FD.t. (12.9.8)


o 0
12.10 The Lagrange principle of stationary action 735

On the other hand,


t t t

J D.Fdt= J (8F + PD.t) dt= J[8F-F(D.t)·]dt+FD.tl~. (12.9.9)


0 0 0

Removing the expression


t t

8 J Fdt = J 8Fdt,
o 0

from these equalities yields the following relationship


t t

D. J Fdt = J [D.F + F (D.tt] dt + (FD.t)t=o· (12.9.10)


o 0

Since
tl tl to

J Fdt = J Fdt - J Fdt,


to 0 0

we immediately obtain
tl h

D. J Fdt = J [D.F + F (D.tr] dt. (12.9.11)


to to

12.10 The Lagrange principle of stationary action


Let us consider a material system subject to holonomic stationary con-
straints. The active forces are assumed to be potential forces so that the
total mechanical energy retains a constant value throughout the motion

T+II = h. (12.10.1)

It is adopted that this relationship remains valid for all neighbouring


paths passing through the two fixed positions q~O) and qF) of the true
path. Since condition (1) imposes a certain restriction on the velocity of the
system particles in the neighbouring motion it would be an error to think
that the system configuration q;
along the neighboring path corresponds
to the configuration qs along the true path at the same time instant. In
particular, we can not require that the passage of the system from its initial
position to the final one takes the same time tl - to as along the true path.
736 12. Variational principles in mechanics

For example, in the case of the free motion of the particle along a straight
line the motion is determined by the equality

(2h
X= v-:;;;:t.
Let Xl denote the position of the particle at instant tl when the particle
moves along the true path. It is not possible to arrive at the same position Xl
moving along the neighbouring path within the same time interval without
changing the energy constant.
Thus, condition (1) requires an asynchronous variation.
For the sake ofthe forthcoming analysis we notice the relationship which
is obtained by varying equality (1) accounting for formula (9.4)

~ (T + II) = 8 (T + II) + (T + IIr ~t = 8 (T + II) = ~h = 8h = o.


(12.10.2)

Let us turn now to Lagrange's fundamental equation (6.4.9), then we


have
d n
dt LPs8qs = 8L = 8T - 8II = 28T.
s=l

Replacing here 8qs by means of eq. (9.3) and noticing that for the stationary
constraints
n
LPsqs = 2T,
s=l

we obtain, due to eq. (9.4), that

d .
28T + dt 2T~t = 28T + 2T ~t + 2T (~tr

2~t + 2T (~t)· . (12.10.3)

Integrating now both sides of the latter equation from to to tll recalling
formula (9.11) and noticing that

(12.10.4)

as the neighbouring paths intersect the true path at time instants to and
tl, we arrive at the equality

J
tl

~ 2Tdt = O. (12.10.5)
to
12.10 The Lagrange principle of stationary action 737

The value

!
tl

A= 2Tdt (12.10.6)
to

is referred to as Lagrange's action. For a single particle we have

! ! !
tl h h

A = mv . vdt = mv . dr = mvds, (12.10.7)


~ ~ ~

where the integration limits to and tl correspond to the initial and final
positions of the particle, respectively. The same expression for A is obtained
for any system of particles if "the velocity vector" v is defined by eq.
(7.8.4), i.e. if the velocity of the representative point is considered on the
Riemannian manifold (7.8.2) for which the square of the arc element is
2T (dt)2. We can also define action A as the sum of the work of momenta
in the true path joining the initial and final positions of the system particles

(12.10.8)

Equality (5) expresses the principle of stationary action: Lagrange's ac-


tion between two fixed positions of the system has a stationary value along
the true path provided that the total mechanical energy retains the same
constant value along all neighbouring paths. This principle was first sug-
gested by Maupertius [64] in a rather obscure form.
In the above reasoning the principle of stationary action was derived from
Lagrange's fundamental equation. But we can adopt another standpoint,
namely we take this principle as the basis statement of dynamics of the
holonomic system subject to stationary constraints and under the action
of potential forces and derive the equations of motion for the system from
it. Then we arrive at Lagrange's isoperimetric problem of the necessary
conditions of a stationarity of functional (6) subject to condition (1). As
mentioned in Sec. 12.4 the matter reduces to construction of differential
equations for extremals of the functional
tl tl

\[1= ![2T+A(T+IT-h)]dt=! Fdt, (12.10.9)


to to

where

F = 2T + A (T + IT - h) (2.10.10)
738 12. Variational principles in mechanics

and)" denotes Lagrange's multiplier.


By virtue of eqs. (9.1) and (9.4) we have

J[~F + J
t, t,

~\[f F (~tt] dt = [8F + P~t + F (~tt] dt


to to

J + J
t, t,

[8F F (~tt] dt = 8Fdt + F~tl~.


to to

The integral is transformed by means of formulae (1.4), (1.8) and (1.9).


Applying also formula (9.3) we obtain

The latter term vanishes due to condition (4). Hence,

(12.10.11)

As (~t)o and (~t)1 are independent and we seek the stationary value of
functional \[f containing Lagrange's multiplier ).., the factors of (~t)o and
(~t)1 are equal to zero, as well as the factor of each 8qs in the integrand.
Then we obtain, first, n Lagrange's equations for F

oC s (F) = .!!:... of
>J
_ of = 0 (S = 1, ... ,n ) (12.10.12)
dt uqs oqs
and, second, the conditions at the end-points

of ) of )
( F- L n
oqs qs = 0, ( F- Ln
oqs qs = O. (12.10.13)
s=1 t=to s=1 t=t,

Let us recall that a consequence of Lagrange's equations is eq. (7.2.4),


which takes the following form

d (F - "-q.
- OF) = -of = )... (T + II -
n
h) = O.
& ~o·
8=1 q8
s at
12.11 Jacobi's principle of stationary action 739

While deriving this result we used equalities (10) and (1). Thus, the value
in parentheses is a constant which is equal to zero at t = to and t = t1,
hence it is held constant for any t

(12.10.14)

Inserting the expression for F into this equation we find

2T + A (T + II - h) - (4T + 2AT) = -2T (1 + A) = O.


Therefore, A = -1 and by virtue of eq. (10) we obtain F = L + h. Hence,
we have proved that equations (12) are Lagrange's equations.

12.11 Jacobi's principle of stationary action


We can avoid the difficulties associated with the necessity of using asyn-
chronous variations if we exclude time from the explicit expression of the
principle of stationary action. This possibility is given by the energy inte-
gral (10.1) which can be written in the form
n n
2T (dt)2 = L L ask (q1,'" ,qn) dqsdqk = 2 (h - II) (dt)2 . (12.11.1)
s=l k=l
This enables dt to be expressed in terms of differentials of the coordinates

n n

s=lk=l
dt = (12.11.2)
2 (h - II)

It is substituted into the expression for Lagrange's action to give

J J
tl (1) n n
A= 2Tdt = 2 (h - II) L L askdqsdqk' (12.11.3)
to (0) s=lk=l

where the limits (0) and (1) correspond to the initial and final positions
of the system. Jacobi suggested to take one of the generalised coordinates,
say q1, as the independent variable. Then A is set as follows

J
qi 1)

A= vRdq1, (12.11.4)
qiO)
740 12. Variational principles in mechanics

and referred to as Jacobi's action whereby function R has the form


n n

s=1k=1

2 (h - II) (all + 2 ~ a1sq~ + ~ ~ aSkq~q~) (12.11.5)

and a prime designates a derivative with respect to q1. The principle of least
action reduces the problem of determining the trajectories for the system to
search of the extremals of functional (4). The differential equations for the
trajectories are the Euler-Lagrange differential equations for the extremals.
There is no need to derive these equations once again since it suffices to
replace L and the independent variable t in the Hamilton-Ostrogradsky
principle by Rand q1, respectively. Then we obtain Jacobi's differential
equations of trajectories for the holonomic system with stationary con-
straints under potential forces

~ ay'R _ ay'R = 0 (s=2, ... ,n). (12.11.6)


dq1 aq~ aqs
The time is determined by quadrature from the energy integral (2)

(12.11. 7)

The solution has 2n constants, namely T, hand 2n - 2 constants due to


integration of eq. (6).
The equations of motion in a more general form can be obtained by
considering the integrand in Lagrange's action (3) as the arc element dlJ of
a certain Riemannian manifold R~. Let us assume

(12.11.8)

and use the notation of tensor calculus, that is the summation sign is
omitted if the index is repeated twice in the summed expression. Then the
values

bsk =2(h-II)ask (s,k=l, ... ,n) (12.11.9)

are the covariant components of the metric tensor in R~. Lagrange's action
is written now, instead of eq. (3), in the form

A = J
(1)
dqS dqk
bsk--dlJ
dlJ dlJ
= J J
(1)

ifJdlJ =
(1)

dlJ = 1J(1) _IJ(O), (12.11.10)


(0) (0) (0)
12.11 Jacobi's principle of stationary action 741

where we have introduced the quadratic form

(12.11.11)

The" arc" (J", taken as the independent variable, represents Lagrange's ac-
tion from initial to the actual point of the trajectory. According to the
principle of stationary action, the true path joining points (0) and (1) of
the manifold R~ is the extremal of functional (10). We can immediately
write down the differential equations by using the above analogy and re-
placing Land t by «> and (J", respectively

~ ~«> 8 - d°«> = 0 (s = 1, . .. , n) . (12.11.12)


d(J" o~ q8
d(J"
They coincide with the differential equations (B.8.12) for geodesics of the
manifold R~ which express the condition that the covariant components
of the curvature vector of R~ along the extremals reduce to zero. In the
contravariant form these equations have the form, see eq. (B.8.7)

d 1q2e> +{ a } d qf3 dq' = 0 (a=l, ... ,n), (12.11.13)


d(J" {3, d(J" d(J"

where Christoffel's symbols are constructed for manifold R with quadratic


form (8).
The independent variable (J" does not appear explicitly in the expression
for «> and this allows us to adopt, without loss of generality, that (J" = 0
and (J" = (J"l in the initial and actual positions of the system, respectively.
In addition to the given constant h, the general integral of system (12) has
2n constants Ck

q8 = q8 ((J", C1 , . .. , C2n; h) (s = 1, . .. , n) . (12.11.14)

To determine the 2n +1 constants C1, ... , C2n; h we have 2n boundary


conditions

( q8)(0) -_ q8 (0 , C1,··· , C2n,. h) , ( q8)(1) =q 8 ( (J"1,C1,···,C2n,. h)


(12.11.15)

and eq. (11) which has no new constants since it is the integral of system
(12). Equation (11) is the normalisation condition which relates the inte-
gration constants of eq. (12). The equation for time, due to eqs. (2) and
(8), is as follows

J
a
d(J"
t = to + 2 (h - II)" (12.11.16)
o
742 12. Variational principles in mechanics

When the forces are absent and n = 2 the manifold R~ becomes R2 which
is a surface in three-dimensional space. The problem of the particle motion
by inertia on the surface reduces to determination of the geodesics of the
surface. The differential equations (12) express a far-reaching generalisation
of this well-known fact. The principle of stationary action is reduced to the
statement that the true path differs from the neighbouring ones in that
the curvature vector along it is zero. The true path is the "most straight"
among the admissible paths which are infinitesimally close to it and have
the same end-points. This form of expression of the principle of stationary
action in the case of motion by inertia was established by Hertz.

12.12 Metric of the element of action and metric of


the kinematic element
We know that the motion of a particle subject to stationary holonomic con-
straint can be related to the motion of representative point on the Rieman-
nian manifold Rn. Its arc element, called by Synge the kinematic element
is determined by the following expression for the kinetic energy

(12.12.1)

When only potential forces are present, the motion of the same material
system is related to the representative point on the Riemannian manifold
R~ with the following metric of the element of the action

(12.12.2)

The trajectories of the representative point in the metric R~ are the geodesics
of this manifold. The relationship between the covariant components of the
metric tensors in R~ and Rn is given by

bsk = 2 (h - II) ask. (12.12.3)

Thus, the determinants of the matrices b and a and their algebraic adjuncts
Bsk and Ask are related in the following way

so that the relationship between the contravariant components of the metric


tensors in R~ and Rn is as follows

bsk 1
= 2(h_II)a s k . (12.12.4)
12.12 Metric of the element of action and metric of the kinematic element 743

The square of the absolute value of vector a in the metric R~ has the
following form

a2 bskasa k = askJ2 (h - II)a S J2 (h - II)a k = askiisiik


bsk asak = a sk as ak - (12125)
= a sk-asak, ..
J2 (h - II) J2 (h - II)
which allows us to relate vector a in R~ to vector ii in Rn due to the rule

(12.12.6)

the factor J2 (h - II) being equal to the value of the" velocity of the point" .
Let us notice that formulae (6) satisfies the relationship between the con-
travariant and covariant components of the vector

The scalar product a . b is invariant since

i.e. the scalar product is the same in R~ and Rn. It follows from this
equation and eq. (5) that the angle between these vectors is the same in
R~ and Rn.
The correspondence of Christoffel's symbols of the first kind in the met-
rics of manifolds R~ and Rn is stated by the formula

~ (8b{3a + 8b'Ya _ 8b{3'Y) (12.12.7)


2 8q'Y 8q{3 8qa

2 (h - II) [p", a] - (a{3aII'Y + a'Ya II{3 + a{3'YIIa),


where the sign '-'"' designates the value in the metric of the kinematic element
and IIa is the derivative of II with respect to qa. The relationship between
the symbols of the second kind is

{pa, }= ba6 [p, ,; 8] = {pa,} - 2 (h ~ II) (a3 II'Y + a~II{3 + a{3'Y IIa ) .
(12.12.8)

Let us consider a certain trajectory of the material system which is a


geodesic of manifold R~ and let T denote the unit vector of the tangent to
the geodesic, then according to eqs. (2) and (6)
1 dqa
(12.12.9)
J2 (h - II) ds
744 12. Variational principles in mechanics

The differential equations (11.12) are equations of the parallel translation


of vector 7 along the trajectory and express the fact that the geodesic
curvature becomes zero. In the covariant notation they have the form

(12.12.10)

The concept of the unit vector of the first normal introduced by formula
(11.15.1) makes no sense for the geodesic, since its curvature k(1) is zero. For
the forthcoming analysis it is necessary to introduce the unit normal vectors
e which are orthogonal to 7 and always remain parallel under translation
along the trajectory. According to Sec. 2.9 the derivative of this vector with
respect to arc of the geodesic vanishes, hence, its contravariant components
satisfy the system of linear differential equations of the parallel translation

(12.12.11)

which is analogous to eq. (10). One particular solution is known, namely


cf3 = T f3 , because equations (10) and (11) coincide in this case. Let b,~
be the vectors which satisfy eq. (11). Its scalar product b.
~ is constant
along the geodesic. This follows from the fact that eq. (11) expresses the
vanishing derivative of e with respect to the arc u of the geodesic, that is
j k
d j k de k
- e· e =-. e + e ·de-
j
= o. (12.12.12)
du du du
For this reason, while considering the system of (n - 1) independent
particular solutions c-h, ... ,c';" of differential equations (11) for ~, ... c
,n 1,
we can subject these solutions to the following initial conditions

(k # j), (12.12.13)
(k = j).

1 n-I
The system of vectors 7, e, ... , e form an orthogonal basis of n unit vec-
tors belonging to manifold R~. These unit vectors move along the trajectory
and do not change their directions, i.e.

d7 = 0 d~
- =0 (k = 1, ... ,n - 1) . (12.12.14)
du ' du
They are determined up to a rotation about 7. If vector p can be repre-
sented on this manifold in the form
1 n-I
p = et(0)7 + et(1) e + ... + et(n-I) e, (12.12.15)
12.12 Metric of the element of action and metric of the kinematic element 745

then, its derivative with respect to 0" is given by

dp I I 1 I n-l
dO" = O:(O)T + 0:(1) e + ... + O:(n-l) e, (12.12.16)

where a prime denotes a derivative with respect to 0". In this regard, this
orthogonal basis replaces the axes of fixed directions.
Using formulae (6)-(8) we can transform eq. (lO) to the form

d T~ ~
\12 (h - II)aa~-d + [(3, /'; O:]T~T'Y -
S J2 (h - II)
1
2 (h _ II) (O:~aII'Y + a'YaII~ - O:~'YIIa) T~T'Y (12.12.17)

a~ dT~ [(3-------:--]- ~ -I'


1 (_ dII ) _
a ds + ,/"O:T T - 2(h-II) TaTs -IIa -0.

Here we used the following relationships

a~'YT
-~-'Y
T = 1, a'Ya II ~T-~-'Y - II ~T-~ -Tag
T --Ta - - radII .r-TaTs'
- _ - dII

(12.12.18)

because II~ are the covariant components of vector grad II. Noticing that

(12.12.19)

and carrying out the analogous transformation of eq. (11) we obtain

(12.12.20)

The vectorial form of eqs. (17) and (20) has the form

dr 1 ( _ dII) de _ grad II . c
ds =- 2 (h _ II) grad II - r ds ' -=r.:::..---.,- (12.12.21)
ds 2 (h - II)'

Recalling Frenet's first formula and taking into account that 2 (h - II) and
grad II are respectively equal to the square of the velocity of the represen-
tative point and the force" acting" on it (see eq. (7.8.6)), respectively, we
can easily recognize that the first equation in (21) is the natural equation
of motion

k(1)n = -1 (- - . TT ) = -Q
Q- Q 1- . nn, (12.12.22)
v2 v2
where nand k(1) denote the first normal to the trajectory and the first
curvature, respectively. Clearly, k(1) -=I=- 0 in the metric of the kinematic
element.
746 12. Variational principles in mechanics

The second equation in (21) indicates that the derivative of vector c


with respect to arc (J" has the direction of T in the metric of the kinematic
element. It is worthwhile stressing that the vector of the first normal in the
metric of the element of the action is no longer translated parallel along
the trajectory.
Let us introduce into consideration a local system of orthogonal axes
1 n-l
T, C, ... , c at point M of the trajectory. Let us consider two infinites-
imally close points Nand N' in the vicinity of M. Let Nand N' lie in
the hyperplanes which are normal to the trajectory at point M and M',
respectively, points M and M' being infinitesimally close to each other.
Denoting

MN=
n-l

LV(k)C
k
and
------jo n-l
MN'=TdS+~(V(k)+dV(k)) c+dc ,
(k k)
k=l
(12.12.23)

we obtain the following formula for the square of the linear element N N
--,

(NN')2 ds" ~ ITds + ~ (V'k)d ~ + ~ dV'k)) I'


2

T {dS + ds ~ V(k) [~~~~~lJ + i dV,k)


where the subscript 0 implies that the value is taken on the supporting
trajectory.
k
Taking into account the orthogonality of the unit vectors T and C, we
obtain

grad II·
[ 2 (h - II)
~l}2 + dV(l)2+ ... + dV(n_l)·
0
2
(12.12.24)

In the particular case n = 3 we introduce the unit vectors ii and b of the


first and second normals to the trajectory, respectively. Since

1 _ 2 _
C = ii cos 1/J + b sin 1/J, C = - ii sin 1/J + b cos 1/J (12.12.25)
12.13 Perturbation of trajectories 747

we can use Frenet's formulae (11.16.1)-(11.16.4) to obtain

;: ~ (~; +k(,,) ~ -k("TOO',", )


(12.12.26)
de
ds =- (d'ljJ
ds + k(2) ) 1-C +k(1)"T_ sm
.
'ljJ.

On the other hand, it follows from eqs. (25) and (22) that

1 1-
2 (h _ II) grad II· c
1 ~ 1 - 1 -
2 (h _ II) grad II· c -Q.
v2
e2 = --Q. nsin 'ljJ = -k(l) sin'ljJ,
v2

and substitution into the second equation in (21) yields


1 2
de k- de k - . n/,
ds = - (l)TCos'ljJ, ds = (l)T sm 'f/. (12.12.27)

A comparison with eq. (26) leads to the relationship between angle 'ljJ and
the second curvature k(2)

J
s
d'ljJ
ds + k(2) = 0, 'ljJ = 'ljJo - k(2)ds. (12.12.28)
o
When the metric of the kinematic element is Euclidean, then, as proved in
1 2
the theory of surfaces, the family of normals e, e form developable surfaces
for which the trajectory under consideration is the edge of regression.
Expression (25) for the linear element referring to the local system of
1 2
axes T, e, e takes the form

[ (h - e1]
grad II· grad II· ~]}2 + dV(l)2 + dV(2)'
2
2 II) 0 + V(2) [ 2 (h - II) 0

(12.12.29)

12.13 Perturbation of trajectories


The problem of stability of the prescribed motion of the material system
can be viewed from a few perspectives. First, we can seek an estimation
of the deviations of the generalised coordinates and generalised velocities
748 12. Variational principles in mechanics

from their prescribed values at any instant provided that the initial pertur-
bations are sufficiently small. This definition of the stability in the sense of
Lyapunov was given in Sec. 11.10 whilst Secs. 11.14-11.17 are devoted to
constructing equations of the perturbed motion which are the variational
equations. Secondly, we can consider only the orbital stability. In this case
the question of time-dependence is relegated to the background and we
study only trajectories of the perturbed motion and establish criteria as
to how close are the trajectories of the perturbed motion to the given tra-
jectory. Lagrange's principle of stationary action turns out to be the most
appropriate for the analysis of the orbital stability since the trajectories of
both prescribed and perturbed motions are the geodesics of the manifold
R~ of the element of the action, i.e. the simplest geometrical objects of this
manifold corresponding to the straight lines in the Euclidean space. This
principle of analysis of the orbital stability is used by Thomson and Tait in
[88] and by Zhukovsky in [97], who additionally studied non-conservative
perturbations.
Basically, the derivation of the variational equations for the perturbed
trajectories repeats that of Secs. 11.14-11.17 with the only difference that
the independent variable is the action along the prescribed trajectory rather
than time t. This action implies the arc a measured along the prescribed
trajectory in the metric of the element of action. The calculation is carried
out in the same metric rather than in the metric of the kinematic element.
_Ct = 0
Let p denote the vector of the perturbation and xC< , x , x designate the
covariant components of this vector, the first and second derivatives of this
vector with respect to arc a respectively. Their expressions are given by
formulae (11.14.9) and (1.14.12)

The variational equations similar to eq. (11.14.17) are

(12.13.2)

where Ry{j{3p denote the covariant components of the Riemann-Christoffel


tensor of curvature on manifold R~.
Let us consider in greater detail the case of the system with two and
three degrees of freedom.
Let n = 2. According to eq. (12.14) it is necessary to put w(1) = ak(l) = 0
in eq. (11.15.27). The result is the following equations

(12.13.3)
12.13 Perturbation of trajectories 749

where € and v are components of the perturbation vector along directions


T and ~. The latter is the only unit vector in R~ which is orthogonal to
T. We limit our consideration to perturbations which are normal to the
trajectory, that is we take € = 0. The perturbation vector then takes the
form
1
p=vc, (12.13.4)

where v denotes "length" in the metric of the element of the action as b is


the unit vector in this metric. In the metric of the kinematic element, i.e.
the surface R2 (but not R2!) with the linear element ds, the length of the
perturbation vector is calculated due to the formula
v v
V= -,,;c2 h=-~II~)
7.'( v'
(12.13.5)

see eq. (12.2).


In eq. (3) the Gaussian curvature of the manifold R2 is denoted by K.
By virtue of eq. (11.15.26) it is equal to

K __ R1212
(12.13.6)
-- b l1 b22 - b~2

Thomson and Tait termed the supporting orbit stable if under a sufficiently
small conservative perturbation the normal deviation v remains bounded
along the whole trajectory. The same definition of the" strength of motion"
was adopted by Zhukovsky. It is worthwhile adding that eq. (3) can yield
only the necessary criterion of the orbital stability. In order to establish the
sufficient criteria we should retain the terms nonlinear in v in the equations
for the perturbed trajectories. The necessary (but not sufficient) condition
for bounds of quantity v, given by eq. (3), is the positiveness of K within
the range of a. This condition is also sufficient for the stationary motions
for which K is held constant along the trajectory.
Let us turn now to the case n = 3. In accordance with eq. (12.14) we
°
should put w(i) = 0,W(2) = in eqs. (11.16.8)-(11.16.10). Under the con-
servative perturbation, i.e. t5h = 0, we obtain the variational equations for
the perturbed trajectories in the form

d:~~l) + K (~, ~) V(l) + K (~,~) V(2) = 0, }


(12.13.7)
2
d V(2)
da 2 +K (2c, 1)
C v(1) +K (2c, 2)
C V(2) = 0,
where v(1) and V(2) are the components of the perturbations vector
1 2
P= C V(1)+ C V(2)· (12.13.8)
750 12. Variational principles in mechanics

Here ~ and ~ denote the unit vectors of the normals translated in parallel
along the supporting trajectory. The coefficients of lI(l) and lI(2) in eq.
(7) are determined due to the rules (11.15.21) and (11.15.23) and are the
quadratic forms of the components of vectors ~ and ~ with the coefficients
defined by Ricci's tensor.
Let us dwell on the case of the Euclidean metric of the kinematic element.
The Cartesian coordinates x, y, z can be taken as qOl, so that

ask=O (s#k), a ss =l (s,k=1,2,3). (12.13.9)

In the metric of the element of the action we have due to eq. (12.7)

[1,1; 1] = -IIx, [1,2; 1] = -IIy, [1,3; 1] = -lIz, [2,2; 1] = II x ,

}
[3,3; 1] = II x , [2,2; 2] = -IIy, [2,3; 2] = -lIz, [2,1; 2] = -IIx,
[3,3; 2] = IIy, [1,1; 2] = IIy, [3,3; 3] = -lIz, [3,1; 3] = -IIx,
[3,2; 3] = -IIy, [1,1; 3] = lIz, [2,2; 3] = lIz,
(12.13.lO)

the other brackets being identically equal to zero.


Calculation of Ricci's tensor by means of these formulae, eqs. (11.15.13)
and (B.14.5) leads to the following values of its contravariant components

ASS = 1 4 [(h - II) ~II + (grad II)2] + B Ss , Ask = B sk ,


8 (h - II)
(12.13.11)

where ~ is the Laplace operator and Bs k denote the contravariant com-


ponents of the auxiliary symmetric tensor of second rank. They are given
by
1
B=- X (12.13.12)
8(h-II)4
3 3
(h - II) IIxx + ~II; (h - II) IIxy + '2IIxIIy (h - II) IIxz + '2IIxIIz
3
(h - II) IIyy + ~II~ (h - II) IIyz + '2IIyIIz
symmetric (h - II) IIzz + ~m
The invariant form of tensor B is as follows

B = _~gradIIgradII _ 1 rad radII (12.13.13)


16 (h - II)4 8 (h - II)4 g g ,

where grad II grad II implies the dyadic product and grad grad II means the
formal dyadic product of the symbolic vector grad and grad II.
12.13 Perturbation of trajectories 751

1 2 . 1 2 .
The covariant components c,\, c,\ of the umt vectors c and c m the coor-
dinate system xC> of the element of the action meet the conditions

1
2 (h - II) L
3 ik
C,\C,\=
{I 0
(i = k)
(i-/=k) ,
(12.13.14)
,\=1

which follow from eqs. (12.4) and (9). By virtue of eqs. (11.15.21) and
(11.15.23) we find

K (c,1c1) = 1
4 (h - II)
3 [(h - II) ~II + (grad II) 2] + 2 2
B,\/l> c,\c/l>'

2 2) =
K ( c,c 1 3 [ (h-II)~II+(gradII) 2] 1 1
+B,\/l> C,\C/l> ,
4 (h - II)
1 2) ,\ 1 2
K ( c,c = - B /l> C'\C/l>.

(12.13.15)

The further transformation is based on change from the Cartesian coor-


dinate system to the local coordinate system of the supporting trajectory.
Using eqs. (13) and (12.6) we can cast the quadratic forms on the right
hand sides of relationships (15) in the form

B,\/l> ~,\~,\= - [ 3
8 (h - II)
3 (~.gradII)2 +
~
--1-----n2 . grad grad II·
4 (h - II)
, ~]
(12.13.16)
1 2 3 1 2
-B,\/l> C,\C,\= 3 c . grad II c . gradII+
8 (h - II)
1 1 2
-------n2 C . grad grad II· c.
4 (h - II)

The matter is reduced to calculating the projections of vector grad II and


the component of tensor grad grad II onto directions of the unit vectors
1 2
C and c. It is adopted that the potential energy, as a function of x, y, z,
is expressed in the vicinity of the supporting trajectory in terms of the
1 2
arc s and the distances ii(l), ii(2) measured along the normals c and c.
The Laplace operator in formulae (15) should be expressed in terms of the
derivatives along these trajectories. All quantities should be determined on
the supporting trajectory. For this reason, having carried out the mentioned
calculation we should take ii(1) = 0 and ii(2) = o.
The Lame coefficients of the linear element (12.29) are

h3=1+ [ ii(l) all ii(2) all ]


--+ --
2 (h - II) aii(1) 2 (h - II) aii(2) 0'
752 12. Variational principles in mechanics

hence
- all 1:. all ~ all k all
gr ad II ='T ha + c
3 S
- -+ c
a v(1) - -,
a v(2) c ·gradll = -a-.
V(k)

Tensor grad grad II is represented in the form of dyadic products

then
~
grad grad II = -a-
a (_' T
all 1:. all ~ all )
a- + C -a- + C -a- ,
C .
V(k)
h3 S Vel) V(2)

1 2
and as the unit vectors T, C, C are independent of V(k) we can conclude that
k j a2 Il
c.gradgradIl.c=a_ a-
V(k) v(j)

The Laplace operator is constructed due to the well-known formulae

~Il = 1 [_a_ h2 h3 all + _a_ h3hl all +..!2. hlh2 all]


hlh2h3 aV(1) hI aV(1) aV(2) h2 aV(2) as h3 as
and, for v(1) = iI(2) = 0, is equal to

a2Il a 2Il a 2Il


[ -2 + a-v(2)
2 + -a 2 +
a v(1) S

1 ( all ) 2 1 ( all ) 2]
2 (h - II) ail(l) + 2 (h - II) ail(2) 0 .

Inserting these formulae into eqs. (15) and (16) we arrive at the expressions

(11) {4 (h -1 II) 2 ( -2-


K c, C = a 2Il + --2
ail
a 2Il) +
as (1)

1
4 (h - Il)3
[32 ( all ) 2 (all) 2] }
ail(1) + as 0

2 C2) = {
K ( c, 1 a 2Il
2 ( -2-
2
a Il) +
+ --2
4 (h - II) ail (2) as

1
4 (h - Il)3
[32 ( all ) 2 (all) 2] }
ail(2) + as 0

K (~ ~) = [1 a 2Il 3 all all]


, 4 (h - Il)2 ail(1)ail(2) + 8 (h - Il)3 ail(l) aV(2) 0
(12.13.17)
12.14 Examples 753

Analogous formulae were derived by Zhukovsky in [97] by means of an


ingenious geometrical construction which is difficult to reproduce.
Note that if we cancel out those expressions in table (10) which do not
contain index 3 we obtain Christoffel's symbols of the first kind for n = 2
provided that the metric of the kinematic element is Euclidean. Due to eqs.
(6) and (B.14.5) we arrive at the formula for Gaussian's curvature

K = { 1 3 [(h - II) (IIxx + IIyy) + II; + II~]} (12.13.18)


4 (h - II) 0

12.14 Examples
12.14.1 Trajectories of a particle under gravity
Dealing with a single particle we can determine the metric of the kinematic
element by means of formulae (13.9). If we use expressions (13.10) for
Christoffel's symbols of the first kind, the differential equation (12.10) for
the trajectory takes the form

2 (h - II) x" - x ,2 IIx + (y12 + Z12) IIx - 2xly'IIy - 2X' zlII z = 0,


(12.14.1)

where a prime denotes the derivative with respect to a. The other two
equations can be written down by analogy. By virtue of eq. (11.11) the
first integral is

(12.14.2)

Using this equation we can easily transform the system of equations (1) to
the form

d ( ) IIx
+ 2 (h _
I
da 2 h - II x II) = 0,

d ( ) IIy
+ 2 (h _
I
da 2 h - II y II) = 0, (12.14.3)
z
d (
da 2 h - II z
) I
+ 2 (hII_ II) = O.

In the case of a homogeneous gravitational field we can restrict our anal-


ysis to planar motion. Directing axis y along the upward vertical we arrive
at two equations

d d ( ) I
da 2 (h - gy) x' = 0, da 2 h - gy h + 2 (h 9_ gy) = O. (12.14.4)
754 12. Variational principles in mechanics

Integrating the first equation and removing y' from the second one by
means of the integral (2), we obtain

X' = 2 (h ~ gy)' d~ EV2 (h - gy) - C2 + 2 (h ~ gy) = 0, (12.14.5)

where E = sign y' = ±1. Using the notation

rJ = EV2 (h - gy) - C2, (12.14.6)

the second equation in (5) takes the form

(12.14.7)

which can be easily integrated. Thus, Lagrange's action is set as follows

(12.14.8)

Here rJo denotes the initial value of rJ


• I
EO = SIgn Yo· (12.14.9)

Let us take yb > 0, then EO = E = 1 on the upward part of the trajectory,


on which y increases from Yo to y*

corresponding to rJo = 0. On the downward part y decreases from Yo to y*


and E = -1.
The constant C can be expressed in terms of the initial value xb of the
derivative x'. Replacing then da in eq. (5) by its value (7) we come, after
integration, to the trajectory equation
X'
X = Xo + 2 (h o - gyo) -.J!. (rJo - rJ) . (12.14.10)
9
Hence, on the upward part

x = Xo + 2(h o - gyo) ~ [)2 (h - gyo) - 4(h - gYO)2 x~-


)2 (h - gy) - 4 (h - gYO)2 Xb2 ] , (12.14.11)

at the turning point of the trajectory

x* = Xo + 2 (h - gyo) xb )2 (h - gyo) - 4 (h - gYO)2 x~ (12.14.12)


9
12.14 Examples 755

and on the downward part

x Xo + 2 (h o - gyo) ~ [J2 (h - gyo) - 4 (h - gYO)2 xb2+

J2 (h - gy) - 4 (h - gYO)2 Xb2] , (12.14.13)

For a given h the solution has three constants, namely Xo, Yo, xb and
describes a bundle of trajectories with the turning point at point xo, Yo.
Any particular curve G of the bundle is specified by slope xb. On a curve
G' of this bundle, which is infinitesimally close to curve G, the slope is
x~ + 8x~ and variation 8x, determining the transition from G' to G, is
given by
ax ,
8x = -a,8xo· (12.14.14)
Xo
From eqs. (11)-(13) we obtain

ax _ x - Xo [ 4 (h - gyo) x~ 1'
'rJo I'rJ I
, - , 1 ± (12.14.15)
axo Xo
where the positive and negative signs correspond to the upward and down-
ward parts, respectively. At the turning point of the trajectory

ax = x* - Xo
aXo"
[1 _ 2(h - gyo) X~2 1 (12.14.16)
Xo 1 - 2 (h - gyo) x~2 •

It follows from eqs. (14) and (15) that variation 8x reduces to zero for
non-zero 8x~ at y = YI which ensures that the square brackets in eq. (15)
vanish. Then by means of eq. (13) we find the corresponding value of x = Xl
and y = YI
2 2

YI = 9h - (h9 - Yo ) 2 (h - gyo) xb
1 _ 2 (h _ gyo) xb2 =
h
9 - ( 9h - Yo ) xb
yb2 '
)

[2 (h - gYO)]3/2 x~ 2 (h - gyo) x~
Xl = Xo +
9 J 1 - 2 (h - gyo) X~2
= Xo +
9
,.
Yo
(12.14.17)
Thus, the infinitesimally close isoenergetic trajectories intersect at point
(Xl, YI) which is the kinetic focus corresponding to the initial point (xo, yo).
As shown in Sec. 12.8 there exist no kinetic foci when paths of the same
duration under motion in the gravitational field are considered.
Expressions for the coordinates for the kinetic focus (17) can be written
in a more transparent form when we notice that
dx dx , i; cos a
x' XO=2 = - - ,
drY v 2 dt 2 (h - gy)' vo vo
756 12. Variational principles in mechanics

FIGURE 12.7.

where Vo denotes the value of the vector of initial velocity and a is the
angle between this vector and axis x. Then we arrive at the formulae

Xl
V2
= Xo + ....Q. cot a, YI
V5
= Yo + 2g (1 - cot a).
2
(12.14.18)
9

Excluding a from these we obtain the equation for the locus of the kinetic
foci corresponding to the initial point

V6 9 (Xl - XO)2 h 9 (Xl - XO)2


~-~=--
2g 2V6
=--~--
9 4 (h - gyo)
. (12.14.19)

This equation describes the parabola of safety which is the envelope of


the family of parabolic trajectories corresponding to the given value of the
kinetic energy and originating at point (xo , Yo). The tangency points of the
trajectory and the safety parabola lie above the level y = Yo of the initial
point for a > 45° and below it for a < 45°, see Fig. 12.7.
We proceed now to examples illustrating the investigated method for
the perturbed trajectories developed in Sec. 12.13. The examples are taken
from the treatise by Zhukovsky [97] where other approaches are applied.

12.14·2 Motion of a particle in central force field


In this case the potential energy is a function only of the distance r from
the centre of the force, hence

X
TIx = TIr -,
r
12.14 Examples 757

By virtue of eq. (13.18) we find

K= 1 3 (h-IT
---dd rITr+ITr 2) . (12.14.20)
4 (h - IT) r r
For instance, we can take
(12.14.21 )

where the factor sign n is entered to ensure that the force is attractive for
fJ > O. Then we have

hfJlnlnrn~2 2hfJlnlnrn~2
K = --'-':"""":"---,,- (12.14.22)
4 (h - IT)3 v6
The necessary condition for stability of the orbit is the positiveness of hn.
In particular, for the circular orbit

so that

the above criterion leads to the inequality n > -2. The case n -1
corresponds to Newton's attraction force, then
fJ
h=--
2r'

and due to eq. (13.3)


a r,;::-;; a
+ Vay rfJ sin
I
v = Va cos rri/ rri/' (12.14.23)
yrfJ yrfJ
It is easy to obtain the dependences versus time

2 fJ t
a=vt=- V = Va cos V[ii r;;. [ii
~t + vay -; sm V~t.
I
(12.14.24)
r '
The perturbation period coincides with the period of the unperturbed orbit
as we could expect.

12.14.3 Motion of a particle on a conical surface


A particle moves on the surface of a cone whose axis comprises angle A
with the downward vertical, see Fig. 12.8. The opening of the cone is 2fJ
and

fJ < A,
758 12. Variational principles in mechanics

z
FIGURE 12.8.

The unperturbed motion begins at the vertex of the cone and occurs on
the lower (OA) and upper (OB) generating lines. The equations for the
perturbed trajectories are required.
The position of point M on the conical surface is described by the dis-
tance l = OM from the vertex measured along the generating line and angle
cp between the vertical plane containing and the plane passing through the
point and the axis of the cone, see Fig. 12.8.
The kinetic and potential energies are

T = -2182 = -21 + l2<p2 sin2 fL) ,


(i2 }
(12.14.25)
II = -gz = -gl (cos fL cos A + sin fL sin Acos cp) .
In this case h = 0 and the square of the element of the action is
dcr 2 = -2IIds 2 = -2II (dl 2 + l2 sin 2 fLdcp2) . (12.14.26)
The metric on the conical surface is Euclidean whereupon formula (13.18)
for the curvature of the manifold of the element of the action is valid.
However the calculation of the Laplace operator and the square of the
gradient should be carried out in curvilinear coordinates l, cp
1 1
~II IIll + -l III + l 2 sm
. 2 II<p<p
fL
-Tcot fL (cos Asin fL - sin Acos fL cos cp) ,

(gradII)2 II2 + 1 II2


I [2 sin 2 fL <p

l [( cos fL cos A+ sin fL sin Acos 'P) 2 + sin 2 Asin2 'P] .


12.14 Examples 759

The value of K should be determined on the supporting trajectories, i.e.


at ip = 0 and ip = Jr. For these trajectories

-11 glcos()..~J-L),

~11 ±TcotJ-Lsin()..~J-L), (grad 11)2 =g2cOS2()..~J-L)

and substituting into eq. (13.18) yields

K sin)..
=±--------~~---
4g[3 sin J-L cos 2 ().. ~ J-L)'
where the upper (lower) sign corresponds to the lower (upper) generating
line of the cone. Differential equation (3) takes the form
d2 // sin).. 0
-± //=
d0"2 4g[3 sin J-L cos 2 ().. ~ J-L)
and it remains to express l in the supporting motion in terms of the inde-
pendent variable 0". By virtue of (26) we obtain

JV-2I1dl=~V29COS()..~J-L)l3/2,
I

0"=
z3 = 90"2
8gcos ().. ~ J-L)
o
and furthermore
d 2 // 2 sin).. //
-± - =0 (12.14.27)
dO"
2 9 cos ().. ~ J-L) sin J-L 0"2 .
We arrive at the equation of Euler's type which is integrable in quadrature.
The solution has the form

1 2 sin)..
"4 ~ "9 -'si-n-J-L-c-o-S7()..-~-J-L-:-)"
The value of // increases beyond all bounds as 0" increases. We could expect
this for the supporting motion along the upper generating line as K < 0
on it. For the lower generating line K > 0 but tends to zero as 0" ---7 00 and
the necessary criterion of stability is also not satisfied.
The value of //, which is the "length" in the metric of the element of
action, is related to the normal deviation

v= l sin J-LDip
from the supporting trajectory by eq. (13.5). Hence
// 2// 2 1
Dip = = . = - . - - (Cw q + C20"~q).
lsinJ-LV2glcos()..~J-L) 30"SlllJ-L 3S111J-Ly(T

On the lower generating line q < ~ and Dip ---7 0 as 0" ---7 00.
760 12. Variational principles in mechanics

12.14.4 Motion on a circle in the field of attraction of two


centres
Motion of a particle on a circle of radius a occurs under the action of
attraction forces of two centres F and Fl. The distance F Fl = 2b and line
F Fl intersects the centre of the circle and is perpendicular to its plane. The
potential energy is given by formula (21). An investigation of the stability
of the circular trajectory is required.
Since the supporting trajectory is a plain curve, angle 'l/J due to eq. (12.28)
1 2
can be taken to be zero. By virtue of eq. (12.25) vectors c and c have
directions of the principal normal to the trajectory and the binormal. In
the case under consideration they are directed to the centre of the circle
and along the perpendicular to its plane. The deviations V(l) and V(2) from
the supporting trajectory are measured along these directions.
The potential energy on the circular trajectory is equal to

n
II = 2{L (a 2 + b2 )"2" sign n

and does not depend upon the position of the particle on the trajectory as
it is independent of 0'. For small deviations

II = {L (rf + r~) sign n = {L { [(a - V(l))2 + (b - V(2))2] ~+


[(a - V(l))2 + (b + V(2))2] ~} sign n
n
2
= {L [ (a + b2 - - -
2aV(l) - 2bv(2)
-2
+ v(l) -2 "2
+ V(2)) +

(a
2
+ b2 - -
2aV(l)
-
+ 2bv(2) -2
+ v(l) -2 "2
+ V(2)) n] SIgn
.
n,

It is sufficient to retain the quadratic terms in the expansion of II in series


in terms of v(l)and V(2). Then we have

II

(12.14.28)
12.15 Rotation of a near vertical rigid body 761

and furthermore

n-4
2f.l (a 2 + b2 )-2- [a 2 (n - 1) + b2 ] Inl,
n-4
2f.l (a 2 + b2 ) -2- [a 2 + (n - 1) b2 ] Inl ,

In order to determine h we construct the equation of motion along the


supporting trajectory

With the help of formulae (13.17) we find the coefficients of the equations
for the perturbed motion

2f.llnl a4 (a 2 + b2 t/ 2 '
a2 + b2 (n - 1)

The motion along the supporting trajectory is stable under the following
conditions

a2
n>l- b2 • (12.14.29)

12.15 Rotation of a near vertical rigid body


Application of the geometry of manifold of the element of the action for
analysis of the perturbed trajectories becomes rather laborious for systems
with number of degrees of freedom n ~ 3. A faster and more reliable way
to construct the differential equations for the perturbed trajectories is to
apply Jacobi's equations (1.6). However the geometrical clarity, as well as
the compactness and elegance of tensor calculus, are lost in this case.
Under notation (11.9), expression (11.5) for R is written in form

(12.15.1)
762 12. Variational principles in mechanics

The generalised coordinates q2, .. . , qn are assumed to be chosen such


that their zero values correspond to the supporting motion. These coordi-
nates and their derivatives q~, . .. , q~ are small along the perturbed trajec-
tories. Hence, only the terms linear in q2, . .. , qn, q~, ... , q~ should be kept
in the differential equations of the perturbed trajectories which are the
variational equations. This allows us to replace coefficients bsk by their ex-
pansion in series in terms of q2, . .. , qn, q~, . . . , q~. The series for bl1 should
be expanded up to the second order terms, the series for bls up to the first
order terms and it is sufficient to take the zero-order terms in the expansion
for bsk when s 2: 2 and k 2: 2.
Within this accuracy we construct the expression for..fR which is needed
for Jacobi's equations (11.6)

VR= ~+ t~q~+~tt (~- blsblk ) q~q~.


s=2 ~ 2 s=2k=2 ~ bl1~ 0
(12.15.2)

The zero indicates that the value in parentheses is calculated for zero values
of q2,··· , qn·
Let us proceed to calculating the derivatives of ..fR appearing in the
Eulerian operator. It is necessary to bear in mind that coefficients bsk
depend upon ql both explicitly and in terms of q2, ... , qn, the latter being
the sought-for functions of ql. Then we have for s = 2, ... , n

d {)R

We arrive at the variational system of linear differential equations

t [( bsk bISb Ik ) " ({) bls {) bIk ) I


k=2 ~- bl1~ 0 qk + {)qk ~ - {)qs Jbil 0 qk

+~ (b~Sk _ blsb Ik )
~ qlk] +~
{) ~
~ _ {)Jbil
{) - 0.
- (12.15.3)
dql v bl1 bl1 bl1 0 ql V bu qs

As an example we consider the case of a rigid body having a fixed point


o and rotating near the vertical. Let A, B, C denote the principal moments
12.15 Rotation of a near vertical rigid body 763

of inertia with respect to axes of the coordinate system Oxyz fixed in the
body. The centre of inertia of the body lies on axis Oz which coincides
with the vertical OC; in the supporting motion. In the perturbed motion
the body position is determined by Euler's angles 'IjJ, {), cp, angle {) being
small. The projections of the angular velocity vector of the body onto the
axis bound to the body are

Wx = ~coscp+~sin{)sincp,wy = -~sincp+~sin{)coscp,wz = (p+~cos{).

Zhukovsky [97] suggested new variables which allow the terms with cos cp
and sin cp to be eliminated in the expression for the kinetic energy

T ="21 (2
Aw x + Bwy2 + CW 2)
z

in the case of small {). Assuming

we can recast the formulae for projections of the angular velocity in the
form

Wx ~coscp + (<i> - (p) {)sincp = ({) cos CPt + <i>{) sin cp,
Wy -~ sin cp + ( <i> - (p) {) cos cp = - ({) sin CPt + <i>{) cos cp,

Wz (P + (<i> _ (p) (1 _ ~2) = ~2 (P + <i> (1 _ ~2) .


Introducing the new variables which are projections of the unit vector of
axis OC; on axes Ox and Oy

x = {)sincp, y = {)coscp,

we obtain
x (P xy - yx
tancp = -,
y cos 2 cp y2

Hence

and the expression for the kinetic energy takes the form

T = ~<i>2 [C+(A-C)X2+(B-C)y2+(2A-C)xy'+
(C - 2B) x'y + Ay,2 + Bx P ] ,
764 12. Variational principles in mechanics

where a prime denotes a derivative with respect to <P. The potential energy
is given by

( {)2)
II = Ql cos {) = Ql 1 -""2 = Ql
( X 2 + y2 )
1- 2 '

where Q and l denote the weight of the body and the abscissa of the centre
of inertia of the body, respectively. From the energy integral we obtain the
value of the constant h in the supporting motion for {) = 0

where <l>o = (wz)o. This value of the energy constant is also held in the case
of conservative perturbation. Then we have

and due to eq. (1) we obtain the following expression for function R

R = [C<l>~+Ql(X2+y2)] [C+(A-C)x2+(B-C)y2+
(2A - C) xy' + (C - 2B) yx' + Ay,2 + BX,2] . (12.15.14)

Therefore, if we ascribe to the coordinates indices 2 and 3, respectively, we


obtain, with the required accuracy, that

(12.15.5)

Canceling out the multiplier <l>o we can set the equations of motion (3)
as follows

Bx" + (C - B - A) y' - (~~ + A - C) x = 0, }


(12.15.6)
Ay" + (C - B - A) x' - (~~ + B - C) Y = o.
Looking for the particular solution in the form

x = M sin ("\<P + a), y = N cos ("\<P + a),


12.16 Hamilton's characteristic function 765

we come to the characteristic equation for .x

( B.x 2 + Ql
<P6 + A - C ) ( A.x 2 +Q
<P6 + B l
- )
C - (C - B - A) 2 .x 2

= AB.x4 - [AB + (A - C) (B - C) - (A + B) ~~] .x2 +

(~~ +A-C) (~~ +B-C) =0.


Hence, the necessary condition for stability of the vertical position of the
axis is positiveness of the roots .xi,.x~ of this equation. This yields the
following criterion

b = AB + (A - C) (B - C) - (A + B) 9 l > 0,
<1>2 o

C = (~~ + A - C) (~~ + B - C) > 0 (12.15.7)

b2 - 4ABc > O.
Let l = 0 which corresponds to the case of rotation about the immovable
centre of inertia. The characteristic equation has the roots
1
.xi = 1, .x 22 = AB (B - C) (A - C) ,

and rotation about the axis of the maximum or minimum moment of inertia
is stable.
For A = Band l > 0 we arrive at the condition for stability of the
vertical spinning top
",2 4AQl
'±' 0 > ----cf2'
This condition also follows from formula (7.9.28) is we take A2 = C~ =0
and use the notation of the present section.

12.16 Hamilton's characteristic function


Let us consider a material system subject to stationary holonomic con-
straints under the actions of potential forces.
In this case the Hamiltonian function does not contain time explicitly and
the principal Hamiltonian function introduced in Sec. 12.8 is the complete
integral of Jacobi's differential equation

~~ +H(Ql, ... 'Qn':~' ... ':!) =0, (12.16.1)


766 12. Variational principles in mechanics

depending on time, the initial and actual values of the generalised coordi-
nates Ct s , qs. Time can appear in the integral of the system of differential
equations of motion only in terms of the argument t - to since these equa-
tions do not contain t explicitly and do not change under the replacement
of t by t - to. Hence,

(12.16.2)

By means of eqs. (10.13.12)-(10.13.14) the complete integral of Jacobi's


equation (1) can be written alternatively in the form

V = - ht + W (q1, ... , qn, "Y 1, ... , "y n-1 , h) + "Yn


where "Yn is an additive constant. After a proper choice of "Yn we obtain

V -h(t-tO)+W(q1,'" ,qn,"Y1,'" , "Y n -1,h) -


W( Ct 1,'" ,Ctn ,"Y1,'" , "Yn-1,h) . (12.16.3)

In order to determine the constants "Y 1, . .. , "Y n-1 and time t - to we can
construct the following n equations with the help of eq. (8.14)

(12.16.4)

From these equations we find the constants "Y 1, ... , "Y n-1' h which, by sub-
stituting into eq. (3), yield the principal Hamiltonian function. But we
can also proceed differently than above. From eq. (4) we can express t -
to, "Y 1, . .. , "Y n-1 in terms of Ct1, . .. , Ct n , h and then exclude them from eq.
(3). The result is

S + h (t - to) = W (q1, . .. , qn, "Y 1, ... , "Y n-1 , h) -


W( Ct1,,,. ,Ctn,"Yu'" , "Y n -1,h) =A(q1,,,. ,qn,Ct1,,,· ,Ctn,h),
(12.16.5)

where S denotes the principal Hamiltonian function.


As follows from eq. (10.13.13), the introduced function A is the complete
integral of the partial differential equation

(12.16.6)

in which the constants Ct1, ... , Ct n denote the initial values of the gener-
alised coordinates.
12.16 Hamilton's characteristic function 767

The function

A (ql, ... ,qn, al,··· ,an, h) = S (ql, ... ,qn, al,··· ,an, t - to) + h (t - to)
(12.16.7)

is referred to as Hamilton's characteristic function. As pointed out in Sec.


12.8 the principal Hamiltonian function S is a special form of Hamilton's
action. By analogy, the characteristic function represents a form of La-
grange's action. Indeed, by virtue of eqs. (8.1) and (8.2) and when the
system possesses the energy integral

T + II = h,

we have

J J J
t t t

S = Ldt = (T - II) dt = 2Tdt - h (t - to) = A - h (t - to).


to to to
(12.16.8)

Casting here Lagrange's action in the form of eq. (11.3) or (11.4)

(12.16.9)

and expressing it in terms of the initial and actual values of the generalised
coordinates we arrive at formula (5), with S being the principal Hamilton's
function.
It can be seen from eq. (9) that we can follow another path to con-
struct Hamilton's characteristic function. On inserting the general solution
of Jacobi's differential equations (11.6)

qs=qs(ql,CI, ... ,C2n-2) (s=2, ... ,n). (12.16.10)

into the following expression

we can express the constants CI, ... ,C2n-2 in terms of ql, ... ,qn, aI, ... ,an·
To this end, we should use both equations (10) and the initial conditions

(12.16.11)

We obtain the properties of the characteristic function which are analo-


gous to those of the principal function, cf. eq. (8.8), by varying the basic
768 12. Variational principles in mechanics

relationship (5). The general case of asynchronous variation is implied in


which the two positions of the system

are compared at different values hand h + {jh of the energy constant,


notation !:lqs being given by formula (9.4).
We obtain

where by virtue of eq. (9.4)

and by analogy

8S n 8S
Ft!:lt + h!:lt + L 7)!:lqs
s=l qs

= L 7){jqs + a + L 7)Qs + h
8S
n

s=l qs
(8S
t
8S
s=l qs
)
n
!:It = Ln 8S
7){jqs + (L + h) !:It.
s=l qs
Relationship (12) takes the form

L -
(8A
n
s=l 8qs
- -
8S) {jqs + L (8A
8qs
- - -
8S ) {jas +
s=l 8as 8as
n

(2T - L - h)!:lt + [8A


8h - (t - to)] {jh = 0
.
The coefficient of !:It disappears by virtue of the energy integral (7). The
first equation in (4) and eq. (5) yield
8A
8h = t - to· (12.16.13)

It remains to equate the coefficients of the independent variables {jqs and


{jas to zero. Then recalling eq. (8.8) we arrive at the following relationships
8A 8S 8A _ 8S - {3
-=-=Ps -- - ~ - - S' (8 = 1, ... ,n), (12.16.14)
8qs 8qs ' 8a s ua s
12.16 Hamilton's characteristic function 769

where Ps and (3s denote the generalised momenta and their initial values.
Hence, similar to the principal function, Hamilton's characteristic func-
tion answers the question as to what initial momenta were applied to the
material system under a prescribed energy constant h, provided that the
system's initial and final positions are given, and what actual momenta the
system has at the final position. Among n equations of the second set in
eq. (14) only n - 1 equations are independent, otherwise the actual values
of all coordinates qI,... ,qn can be expressed in terms of the constants
aI,'" ,an'},'" ,(3n,h. However it is not possible since these constants
enable the n - 1 coordinates q2,... ,qn to be expressed in terms of qI.
The characteristic function contains the integral of the system (11.6) of
differential equations of trajectories rather than the differential equations
of motion.
Recalling that momenta Ps are the covariant components of the velocity
vector v of the representative point in the Riemannian space with the met-
ric of the kinematic element 2Tdt 2 , we can write the first set of equations
in eq. (14) in the form

v = gradA.

Let us consider now the case of a free particle. At the initial position it
possesses velocity whose direction is immaterial and whose value is deter-
mined by the energy constant h. We can speak about not a single particle
but about an infinite number of identical particles thrown in all possible
directions. All these particles reach (but not simultaneously) the surface
A, the velocity of each particle being normal to this surface and described
by the coordinate (x, y, z) on the surface A. While carrying out dynamical
investigations Hamilton was guided by the optical analogy for which sur-
faces A = const are the wave surfaces (on which t - to = const) and the
particle trajectories are the trajectories of a beam of light which are normal
to the wave surfaces. The principle of the stationary action is related to
Fermat's principle in geometrical optics which expresses the requirement
of stationarity of the following integral

J
(2)

n(x,y,z)ds,
(1)

where n denotes the refractive index of the inhomogeneous isotropic medium


under consideration. Comparison with eq. (10.7) shows that the corre-
sponding dynamical quantity is the value of the momentum

mv = J2m [h - II (x, y, z)].


In other words, the trajectory of the light beam coincides with the trajec-
tory of the material point moving in the field with the following potential
770 12. Variational principles in mechanics

energy

n2
II=h--.
2m
Details of the optical analogy that played an important part in the devel-
opment of wave mechanics can be found for example in [95] and [49].
Let us turn to examples of constructing the characteristic function.

12.16.1 Motion in a gravitational field


In the problem of motion of a particle in a homogeneous gravitational field
the action is expressed by eq. (14.8)

(Y = ~ (7]0 - 7]) [~ (172 + 7]017 + 7]~) + C 2 ] , (12.16.15)

where, due to eqs. (14.6), (14.9) and (14.10),


. C
x - Xo = 9 (7]0 -7]) , 170 = y'2 (h - gyo) - C2. }
(12.16.16)
7] = y'2 (h ....:-gyo) - C2.

The characteristic function is obtained by removing C 2 from expression


(15) as in this case the action is expressed in terms of the coordinates of
the initial (xo, Yo) and the final (Xl, Yl) positions of the particle and the
energy constant h.
The calculation is as follows, cf. [85]. We have

2 (y - Yo) C
170 + 17 = ----'-'---'--'---
X-Xo

and we use the identity

to come to the following equation for C 2

(12.16.17)

Denoting for brevity

(2.16.18)
12.16 Hamilton's characteristic function 771

we obtain from eq. (17)

02 = M ± VM2 - N2 = _1_ (V M + N ± VM _ N) 2 }
g2 (x _ XO)2 2N2 4N2 '
g2(x;;XO)2 =2(M~VM2_N2) = (VM+N~VM-N)2.
(12.16.19)

Casting now eq. (15) in the form

A=a = ~ (1]0 -1]) [~(1]0 + 1])2 + 112 (1]0 _1])2 + 0 2 ]


= X - Xo [( y - Yo ) 2 0 2 + ~ g2 (x - XO) 2 + 0 2]
o X - Xo 12 02

= ~ ( VM + N ~ VM - N) [~ (M ± vi M2 - N2) +

~ ( M ~ vi M2 - N2) ]

= 3~ ( VM + N ~ VM - N) (2M ± vi M2 - N2) ,

we come to the following expression for the characteristic function

At the initial point of the trajectory N = 0, taking the plus sign in the
latter equation, we have A = O. The plus sign is kept unless N = M. It is
easy to prove by means of formulae (14.17) that this occurs at the kinetic
focus. Hence unless the kinetic focus is reached

(12.16.20)

and after this

(12.16.21 )

In accordance with eq. (13) we have

(12.16.22)

This is the equation of the curve which contains all the particles emanating
from the initial point in all possible directions with the same initial velocity.
772 12. Variational principles in mechanics

By virtue of eq. (14) we find the initial momenta

8A =..!L [JM+N±JM-N](x-xo)=-xo,
8xo 2N
8A 1 (12.16.23)
- = - - [JM +N=fJM -N] +
8yo 2
2~ [JM + N ± JM - N] (y - Yo) = -Yo,
Each of these equations represents the equation of a bundle of the isoen-
ergetic trajectories originating at point (xo, Yo). Each curve is specified by
the value Xo or Yo of the slope. Noticing that

_1_ =..!L (JM +N±JM - N)


t - to 2N
we arrive at the following notation of equalities (23)
x-xo . 1 y - Yo .
-
t -- to
- -- x0, - g (t - to)
2
+- -=
t - to
Yo·

12.16.2 Keplerian motion


In the problem of Keplerian elliptic motion, Sec. 10.15, integral (10.14.6)
of Jacobi's partial differential equation reducing to zero at the initial point
(ro, <Po) of the trajectory has the form

J
T

W = f3<.p (<p - <Po) + V2hr2 + 2p,r - f3~ d;, (12.16.24)


TO

and equality (4) is as follows

8W
8f3<.p = <P - <Po - f3<.p
J T

rV2hr2
dr
+ 211r _ 132 = o.
(12.16.25)
TO ,- <.p
This is the equation for the bundle of isoenergetic trajectories originating
from point (ro, <Po). Using this in expression (24) leads to the form

W = J
T

TO
2hr + 2p,
V2hr2 + 2p,r - f3~
dr. (12.16.26)

The characteristic function should be obtained by removing (3<.p from


equalities (25) and (26). In order to simplify the further equations let us
introduce the notation

(12.16.27)
12.16 Hamilton's characteristic function 773

where a, e and p denote the major semi-axis, the eccentricity and the pa-
rameter of the elliptic orbit, respectively.
Then we can write

Introducing the notation

y'2ar - r2 - a2 (1 - e2) = (, V2aro - r5 - a2 (1 - e2) = (0 (12.16.28)

we arrive at the equalities

e2rro cos (ip - ipo) = (1 - e2) ((0 + (a - r - ae 2) (a - ro - ae 2),


(12.16.29)

W ( -_
y'aIL { _ 1 [((0 + (a - r) (a - ro)] } .
(0 - arccos 22
= (12.16.30)
a a e

The identities

a (1- e2) - r = Vr 2e2 - (1- e2) (2, }


(12.16.31 )
a (1 - e2) - ro = Vr5e2 - (1 - e2) (6,

follow from formulae (28). By using these, eq. (29) has the form

Taking the square of both sides, cancelling out the factor e 2 and applying
identities (31) we come to the relationship

e2r2r5 [1 + cos 2 (ip - ipo)] - 2rro cos (ip - ipo) [a (1 - e2) - r] x


[a (1 - e2) - ro] - (1 - e2) (r2(~ + r5(2) = 0, (12.16.32)

which is a quadratic equation for p. It determines the value of p on the


ellipse of the bundle passing through the points (ro, ipo) and (r, ip)

1 1 2
F(p)= [ 2"+2--cos(ip-ipo) ] p2 +
r ro rro
[ l+COS(ip-ipo)
a -2 ~+ro
(1 1)] [l-cos(ip-ipo)]P+
[1 - cos (ip - ipO)]2 = O. (12.16.33)
774 12. Variational principles in mechanics

We find two solutions

--:-_l-:--_c_o--:s,.o.:('P'-----..,.:'P--"O:..:. .)_..,- [_ 1 + cos ('P - 'Po) + ~ + ~ ±


PI,2 =
-1 + -1 - -----'.'--~"'-
2 cos ('P - 'Po) 2a T TO
T2 T5 TTO
r-l-+-CO-S-(-'P---'P-O-)-_--I-_-_-I~-+-2~l
viI + cos (cp - 'Po) (12.16.34)
4a 2 aT aTo TTO '

corresponding to two curves joining these points. At the culminating point


(TI' 'PI) of the bundle, i.e. where the curves merge together, the square root
term vanishes and quadratic equation (33) has a double root. This defines
the locus of the kinetic foci of the isoenergetic trajectories conjugated to
the culminating point. Its equation is as follows

4a (2a - TO)
4a - TO
TI = TO (12.16.35)
1- cos ('PI - 'Po)
4a - TO
As TO < 2a and
TO
0< E = < 1,
4a - TO
curve (35) is an ellipse and E is its eccentricity. One of the foci lies in the
attracting centre whereas the second one is spaced at the distance

PI 4a (2a - TO)
2aE = 21E -2 ' PI = ,
-E 4a - TO
where a and PI denote the major semi-axis and the parameter, respectively.
Inserting the expression for E we find that the focal distance equals 2aE =
TO, the second focus of ellipse (35) is located at the initial point and the
major axis 2a is equal to
TO
2a = - = 4a - TO. (12.16.36)
E

The elliptic trajectory and ellipse (35) are marked in Fig. 12.9 by C and
E, respectively.
Let FI and F2 be the foci of the elliptic trajectory and M* be the kinetic
focus conjugated to the initial point Mo. Due to the property of ellipse

so that

(12.16.37)
12.16 Hamilton's characteristic function 775

FIGURE 12.9.

On the other hand, due to the property of ellipse (35) with foci FI and Mo
we have

M*FI + 1'vf*Mo = rl + M*Mo = 2a, M*Mo = 2a - rl = 4a - ro - rl,


and a comparison with eq. (37) yields
MoF2 + F2M* = M*Mo·
If follows from this result that line segment MoM* joining the conjugate
kinetic foci Mo and M* passes through the second focus F2 of the elliptic
trajectory which was pointed out by Jacobi [44].
It remains to notice that when we insert one of the values of p due to eq.
(34) into eq. (30) and replacing e2 by 1 - E we arrive at the expression for
a
Hamilton's characteristic function A which is used unless point M reaches
the conjugate kinetic focus. After this , p is replaced by its second value.
In order to calculate the time in terms of the characteristic function we
use the relationship (13)
oA oA2a2
t-to= - =-- (12.16.38)
oh f.toa
and to determine the actual and initial values of momenta we apply for-
mulae (14)
oA oA
Pr = or' (3r = - oro· (12.16.39)
776 12. Variational principles in mechanics

12.17 On the character of the extremum of


Lagrange's action
As established in Sec. 12.10 Lagrange's action between two fixed positions
of a system has a stationary value along the true path provided that the
neighbouring paths have the same value of the energy constant h. Is this
stationary value a minimum? Let the motion of the system be related to
motion of the representative point on manifold R~ with the metric of the
element of the action. Then the affirmative answer to the question posed
above is given by the following theorem: the geodesic joining two sufficiently
close points of this manifold is shorter than any other line joining these
points.
The theorem is proved as follows. Let us fix the initial position of the
system (qS)o and consider a family of hypersurfaces A (qS, qQ, h), Hamil-
ton's characteristic function being constant on each of these surfaces. All
trajectories belonging to the set oon-l trajectories of the bundle with ver-
tex at the initial position (qS)o intersect orthogonally each hypersurface
at each point. Let us consider the points of intersection of the trajecto-
ries with a certain hypersurface A = Ao and layoff the segments of the
length ao from the intersection points. The loci of the ends of tbe seg-
ments are the hypersurfaces of the same family since the segment of the
arc of geodesic R~ is equal to the change in Lagrange's action. Thereby
a system of "parallel" hypersurfaces and the curves normal to them are
constructed in a small vicinity of the bundle vertex. The position of the
point on the hypersurface which is the (n - I)-dimensional manifold is de-
scribed by (n - 1) parameter xl, ... ,xn - l . We can introduce the system of
coordinates Xl, ... ,xn - l , xn = a where a denotes the arc measured along
the trajectory from the bundle vertex and determines the position of any
other point of manifold R~. Let the arc element of this manifold be given
by the quadratic form
n-l n-ln-l
ds 2 = gskdxSdxk = gnndxndxn + 2 L gindxidxn + L L gskdxsdxk.
i=l 8=1 k=l
(12.17.1)

On the hypersurface A = const, i.e. where dx l = ... = dx n - l = 0 we have


ds 2 = da 2 = gnndxndxn
and thus

However
gin = 0 (i = 1, ... ,n - 1), (12.17.3)
12.17 On the character of the extremum of Lagrange's action 777

which follows from the orthogonality of the trajectory tangent T to any


infinitesimal vector Dr on the surface A = const passing through the point
under consideration and the definition gin as the scalar product of the basis
vectors and T. Hence, on an arbitrary curve of manifold R~ we have
n-1n-1
ds 2 = da 2 + L LgskdxSdxk, (12.17.4)
s=l k=l
where
n-1n-1
L L gsk dxS dx k
s=l k=l
is the square of the arc element on the surface A = const and therefore is a
positive definite quadratic form in differentials dxs. The length of the arc
of any curve of the manifold is given by

J J ~ Jda = a(l) -
(1) (1) n-1 n-1 1/2 (1)

S(l) - S(O) = ds = [da 2 +L L 9SkdXSdx k ] a(O)


(0) (0) s=l k=l (0)

and thus the geodesic is proved to be the shortest distance between two
points in the vicinity of any initial point.
The locus of the kinetic foci conjugated to the origin of the bundle of
trajectories under consideration is a focal surface which is conjugated to this
origin. For example, in the case of motion of the particle in the gravitational
field the safety parabola (14.19) serves as this "surface" whilst in the case
of the elliptic Keplerian motion the ellipse (16.35) plays the role of this
"surface". The position of this focal surface determines the size of the
"sufficiently small vicinity" mentioned above. The boundary is determined
by that surface of the family A = const which contains the kinetic focus
nearest to the origin. There is no need to prove that Lagrange's action
along the trajectory joining the initial and final positions is not a minimum
because the proof would repeat the content of Sec. 12.3 illustrated by Fig.
12.2.
The question of the character of the extremum of Lagrange's action along
the chosen trajectory Co of the bundle is related to the problem of the
stability of this trajectory. Let us restrict our analysis to the case n = 2
and consider the family of the integrals of differential equations for the
perturbed trajectories (12.13.3)

d2 v
da 2 +Kv=O, (12.17.5)

reducing to zero at a = ao. The solution of differential equation (5) is called


non-oscillating in the interval (aI, a2) if it reduces to zero not more than
778 12. Variational principles in mechanics

once. By virtue of the theorem which states that if

(12.17.6)

then the solution of differential equation (5) is non-oscillating in any subin-


terval (O"i, 0";) ofthe interval (0"1,0"2). This means that there exists no per-
turbed trajectory C' which is infinitesimally close to Co, begins at point
0" = 0"0 (0"1 < 0"0 < 0"2) and intersects C within this interval. Thus, this
interval contains no kinetic foci and Lagrange's action along the trajectory
(O"i, 0";) is minimum.
For example, let K = _m 2 = const, then the boundaries 0"1 and 0"2 move
to -00 and 00, and the interval (O"i, 0";) without kinetic foci can be chosen
such that within this interval the absolute value of the solution
v'
v = -...2. sinh (0" - 0"0) (12.17.7)
m
can be made as small as is wished. This can always be realised by a proper
choice of v~. This ensures that C' remains a trajectory which is adjacent
to C and does not intersect C since function (7) does not vanish at 0" 1= 0"0.
Let us assume that

(12.17.8)

Sturm's theorem, [82], states that the distance between two neighbouring
roots of the solution of differential equation (5) is smaller than Jr / v' Kmin
and greater than Jr / v' Kmax. If we take the solution which turns to zero
at 0" = 0"0 we can assert that interval (0"1,0"2) of trajectory Co has the
conjugate kinetic focus provided that
Jr
(12.17.9)

This, along with the definition of the orbital stability in the sense of
Thomson and Tait, explains the assertion of Zhukovsky [97] which is "the
action between two arbitrary positions is smallest for any weak motion and
does not possess this property for the strong motion" .
Returning to the example of motion in the gravitational field we recall
that the energy integral, in the notation of Sec. 12.14, has the form

v2
h= 2" +gy,
hence,

(12.17.10)

where the equality sign corresponds to the vertical motion at the culminat-
ing point.
12.17 On the character of the extremum of Lagrange's action 779

In accordance with the above remark we can find the minimum distance
between the vertex of the trajectory bundle and the safety parabola (14.19),
i.e. the minimum of the quantity

Taking into account inequality (10) it is easy to conclude that the required
minimum is reached at Y = YI corresponding to the equality sign in in-
equality (10). It follows from eq. (14.19) that Xl = Xo. Hence, the point
of the safety parabola that is nearest to the origin lies at the intersection
with the vertical passing through the origin.
Inserting the values
h
X = Xo, Y= - = YI (12.17.12)
g
into formula (16.20), we find the constant in the equation for the curve
A = const

1 { [2h - g (YI
3g + Yo) + g IYI - Yoll 3/2
- [2h - g (YI + Yo) - g IYI - Yol]3/2} = 3~ [2 (h - gYO)]3/2.

The obtained curve

A = 3~ {[2h - g (y + yo) + gJ(x - xO)2 + (y - yo)2f/2

- [2h - g (YI + Yo) - gJ (x - xO)2 + (y - YO)2 f/2}

~ [2 (h - gYO)]3/2 (12.17.13)
3g
is depicted in Fig. 12.10, cf. [93]. It is a closed curve and Lagrange's action
has a minimum along any trajectory with the final point in domain G
bounded by this curve. All curves of the family

(12.17.14)

are also closed and lie in domain G. They contract to the bundle origin
as 'Y decreases, and the "curve" 'Y = 0 degenerates to point Xo, Yo. The
trajectories of the bundle are orthogonal to these curves.
780 12. Variational principles in mechanics

FIGURE 12.10.

Let us consider now the curves A = 'Y for 'Y > 'Y *. Let us fix point (xi, Yi)
of the safety parabola, then making use of eqs. (14.19) we find that

1 3/2
'Y = 3g [4h - 2g (y; + Yo)] > Y*·

The required equations have the form

A = 3~ {[2h-9(y+yo)+gV(X-XO)2+(Y-YO)2f/2 =f
[2h - 9 (Yl + Yo) - gV(x _ xO)2 + (y _ YO)2] 3/2}

~ [4h - 2g (y; + YO)]3/2 . (12.17.15)


3g
The upper and lower signs correspond to the branches Al and A2 shown
in Fig. 12.10. This figure displays two trajectories C* tangent to the safety
parabola at points (±xi, Yi). Let us agree to categorise the bundle tra-
jectories as being "steep" and "flat". Any point of the domain bounded
by curves C* and Al can be connected by a flat trajectory, for which La-
grange's action is minimum. However any point of the domain bounded by
the safety parabola and curve Al is reached by a high trajectory orthogonal
to branch A 2 . The action along the path of this trajectory containing the
kinetic focus is no longer minimum.
Appendix A
Elements of the theory of matrices

We restrict ourselves to the elementary definitions and concepts which are


relevant to the content of the present book. A great number of books are
devoted to matrix theory, see for example [63], [26], [27] etc. In [15] a
detailed matrix calculus is provided and forms an essential prerequisite
to the understanding the equations of analytical mechanics and vibration
theory.

A.I Definitions
An array of m x n values cast in the form of a rectangular table
all al2 al n
a21 a22 a2n

which contains m rows and n columns is referred to as a m x n rectangu-


lar matrix. The first and the second subscripts of the matrix element aik
implies the number of the rows and columns, respectively. In what follows,
the matrix of elements is denoted by the single letter a. This indicates that
the whole array of these elements is under consideration. Matrix theory
suggests the rules of operation on such arrays.
In order to stress that we deal with the entire matrix aik rather than
with single elements considered separately we use the following notation
782 Appendix A. Elements of the theory of matrices

for the matrix


all al2 al n
a21 a22 a2n
a= (A.1.1)

In order to save space, we will also write

a=llaikll, (i=l, ... ,m; k=l, ... ,n). (A.1.2)

Two m x n matrices are said to be equal if their corresponding elements


are equal, that is

a=b if aik=bik (i=l, ... ,m; k=l, ... ,n). (A.1.3)

The matrix is called the null matrix and is denoted as 0 if all of its
elements are equal to zero.
The matrix for which the number of rows coincides with that of the
columns, i.e. M = N, is termed the square matrix of order N.
Let us consider some examples.
1. Let us take an n-dimensional vector x which is given by n values
Xl, X2, . .. , Xn . It can be viewed as a column and cast in the form of a n x 1
matrix

X= (A.1.4)

or as a row which can be understood as a 1 x n matrix

(A.1.5)

2. Having a system of m functions of n variables XI, ••. , Xn

iT = iT (Xl, ... , Xn) (r = 1, ... , m) , (A.1.6)

we can define the m x n matrix

~~ = I ~~: II, (r = 1, ... , m; s = 1, ... , n) (A.1.7)

which is referred to here as Jacobi's matrix of this system of functions.


Particularly, if

fT = 8!f> (r = 1, ... ,n)


8x T
A.I Definitions 783

then we obtain a square matrix


8 2 <f>
H = (r,8 = 1, ... ,n) (A.1.8)
8x s 8x r
referred to as the H e88 matrix of function <f>.
3. The array, each entry of which is a Kronecker's delta

(i i= k)
(i,k=l, ... ,n), (A.1.9)
(i = k)

forms the square unit matrix denoted by En (or simply E when the order
of the unit matrix is clear)

E=118ik l (i,k=l, ... ,n). (A.1.10)

The diagonal elements of En are equal to unity whilst the nondiagonal


elements are zeros.
Returning to the general definitions, let us consider a 8 x t matrix which
is composed of the elements of 8 rows and t columns of matrix a

ai,k, ai,k 2 ai, k t


ai2k, ai2k2 ai2kt
(A.1.11)

aisk, aisk2 aiskt

where iI, i 2 , ... , is are any of the numbers 1,2, ... , m and k l , k 2 ,··· , k t
are any of the numbers 1,2, ... , n. Clearly, 8 ::; m, t ::; n. Then, the 8 x t
matrix (11) is referred to as the submatrix of a and is denoted as

(A. 1.12)
If we deal with a quadratic submatrix, then 8 = t. The determinant,
constructed from the elements of sub matrix (11), is referred to as the minor
determinant of order 8 of matrix a. Clearly, the highest order of the minor
determinant of a m x n matrix is equal to the smallest of the numbers m, n.
Let a be a square n x n matrix. The determinant constructed from the
elements of this matrix is called the determinant of matrix a and is denoted
lal or det (a)
an al2 al n
a21 an a2n
lal = det (a) = (A. 1.13)

The determinant of the quadratic n x n matrix is one of the minor de-


terminants of n - th order, namely that the determinant for which the
784 Appendix A. Elements of the theory of matrices

numbers iI, ... , in and kI, ... , k n coincide with the numbers of the natural
series 1,2, ... , n. It is evident that the other minor determinants of order n
can differ from the above determinant only by sign.
The square matrix whose determinant is not equal to zero is referred to
as non-singular, otherwise it is called singular.
Returning to the case of a m x n matrix, let us denote the highest order
of non-trivial minor determinant as r. We can introduce the concept of
the matrix deficiency, which is the difference between the smallest of the
numbers m, nand r. For example, if m ~ n and the determinants of all m x
m submatrices Iiallmm are equal to zero whereas there exists an (m - 1) x
(m - 1) submatrix Ilall m- 1 m-l with non-zero determinant, then the matrix
deficiency of a is 1. The d~ficiency is equal to zero if there is a non-trivial
minor determinant of order m (for m ~ n) or order n (for m ~ n).
4. Let us consider Jacobi's matrix (7) with m < n and zero deficiency.
Then, there exists a non-trivial Jacobian
8ft 8ft 8ft
8Xk1 8Xk2 8Xk""
D (fI, ... , 1m)
812 812 812
8Xkl 8Xk2 8Xk", =i= 0, (A.1.14)
D(Xk1,···,Xk".)
81m 81m 81m
8Xk1 8Xk2 8Xk""

which enables values Xkll ... , Xk"" to be expressed in terms of the remaining
n - m values Xk",+ll ... , xknfrom the following system of equations

Ir (XI, X2, ... , xn) = 0 (r = 1, ... , m) . (A.1.15)

Let us consider some vertical and horizontal" separators" through matrix


a. These lines can be viewed as borders of the submatrices whose indices
are arranged as numbers in the natural series. The notational shorthand
for these submatrices is as follows
ai+l,k+l ai+l,k+t

II a (i+l,k+l)
ai+2,k+l ai+2,k+t
=
11
(s,t) (A.1.16)

Then we can represent matrix a in the following way

a= (A.1.17)

Up until now we have considered matrices whose elements are real or


complex numbers. However it is worthwhile to study block matrices whose
A.I Definitions 785

elements are matrices, too. For example, if we remove the lines in eq. (17)
then we arrive at the complex square matrix of order 2
(1,1)
II a II (s,t) Il all(1,t+1)
(s,n-t)
(A.l.18)
Il all(s+1,l)
(m-s,t)
Ii all (m-s,n-t)
(s+1,t+1)

The operations on this block matrix are rather similar to the operations
on matrix a.
5. The following expression

[ (3] = ~ (Oqr aPr _ oqr apr) = - [(3 ] (A.l.19)


a, ~ oa 0(3 0(3 oa ,a
r=l

is referred to as Lagrange's brackets of the system of functions qr, Pr with


respect to variables a, (3. It is clear that
[a,a] =0. (A.l.20)
Let us consider now the system of 2n functions qr, Pr of 2n variables
ar, (3r
qr:qr(a1, ... ,an;.(31, ... ,(3n), } (A.l.21)
Pr - Pr (aI, ... , an, (31, ... , (3n),
and construct the square 2n x 2n Lagrange's matrix

[aI, a1] [aI, an] [aI, (31] [aI, (3n]


[a2,a1] [a2,a n ] [a2, (31] [a2, (3n]

[a n ,a1] [an, an] [an' (31] [an, (3n]


A=
[(31' a1] [(31' an] [(31' (31] [(31' (3n]
[(32' a1] [(32' an] [(32' (31] [(32' (3n]

[(3n, a1] [(3n, an] [(3n, (31] [(3m (3n]


(A.l.22)
It can be represented by four n x n matrices denoted by [[aall , [[a(3ll , [[(3all
and [[(3(3ll

A-II
- [[aall
[[(3all
[[a(3ll
[[(3(3ll
I. (A.l.23)

If we assume that the Jacobian is not equal to zero


D (q1, ... , qn;P1, ···,Pn) I- (A.l.24)
D(a1, ... ,an ;(31, ... ,(3n) 0,
786 Appendix A. Elements of the theory of matrices

then system (21) is resolvable for ar, (3r

ar = a r (ql, ... , qn; PI, ···,Pn), (A.1.25)


(r=l, ... ,n).}
(3r = (3r (ql, ... , qn;PI, ... , Pn) ,

The following expression

(A.1.26)

is called Poisson's brackets of the system of functions a, (3 with respect to


variables qr,Pr' As follows from the definition

(a,a) =0. (A.1.27)

The Poisson matrix is introduced by analogy with Lagrange's matrix.


Its elements are Poisson's brackets (ai, ak), (ai, (3k), ((3i' ak), ((3i' (3k)' The
notation analogous to eq. (23) has the form

p -II
-
((aa))
(((3a))
(( a(3))
(((3(3))
I. (A.1.28)

A.2 Operations on matrices


1. Transposing. Given a m x n matrix, we can determine a n x m matrix a'
whose rows and columns are respectively the columns and rows of matrix
a. Then a' is referred to as the transpose of a. Clearly, transposing the
transposed matrix a' yields a

(a')' = a.

Elements aii of the square matrix is called the diagonal. A square matrix
a is said to be symmetric if aik = aki. It is clear that a symmetric matrix
is equal to its transpose

a = a'. (A.2.1)

Hess's matrix (1.8) can serve as an example of a symmetric matrix.


A square matrix a is termed skew-symmetric if aik = -aki. The exam-
ples are Lagrange's matrix (1.22) and Poisson's matrix (1.28). Matrices
[[aa]] , [[(3(3]] or ((aa)) , (((3(3)) are skew-symmetric also.
By virtue of the property of determinants

lal = la'l· (A.2.2)


A.2 Operations on matrices 787

Example 1. To any vector a with projections aI, a2, a3 we can relate the
3 x 3 skew-symmetric matrix

o
a= (A.2.3)

which is used for matrix notation of vector operations.


Example 2. The transpose of a n x 1 matrix-column x is the 1 x n matrix-
row x'.
2. Addition and subtraction of matrices. These operations are defined
for matrices of the same order m x n. The sum and the difference of two
such matrices a and b is the matrix C whose elements are equal to the sum
and the difference of the corresponding elements of the matrices a and b,
respectively, i.e.

c=a±b if cik=aik±bik (i=l, ... ,m; k=l, ... ,n). (A.2.4)

It is evident that

a + b = b + a, a + (b + c) = (a + b) + c = a + b + c.
The definition of the sum and the difference is generalised to the block
matrices. However it is required that the orders of the added or subtracted
submatrices coincide.
3. Multiplication by a scalar. The product of matrix a with a scalar), is
the matrix with the elements equal to ),aik. The notation is as follows

),a=ll),aikll (i=l, ... ,m; k=l, ... ,n). (A.2.5)

In particular, for)' = -1 we have the matrix -a with the elements -aik.


Hence, the transpose of the skew-symmetric matrix yields

a' = -a. (A.2.6)

For instance,

a-I = -a.
-
(A.2.7)

In accordance with the above definitions, the identities

can be cast in the form

a = ~ (a + a' ) + ~ (a - a' ) (A.2.8)


2 2
788 Appendix A. Elements of the theory of matrices

and determine the decomposition of the square n x n matrix a into a sym-


metric part

1( ')' 1 (' ) (A.2.9)


2 a + a =2 a + a

and a skew-symmetric part

~ (a - a')' = -~ (a - a') = ~ (a' - a). (A.2.1O)


2 2 2
Example 3. By virtue of eqs. (1.20) and (1.22), the transpose of submatrix
[[a,Bll of Lagrange's matrix (1.23) is given by

[[a,BjJ' = - [[,Ball· (A.2.11)

The analogous formula takes place for the submatrix of Poisson's matrix
(1.28)

((a,B))' = - ((,Ba)).

4. Multiplication of matrices. Let us consider a m x p matrix a and a p x n


matrix b. The number of columns in a is equal to the number of rows in b.
Such matrices taken in the subsequence a, b are referred to as the conform
matrices. The multiplication is determined for the conform matrices. The
product c = ab is the m x n matrix whose elements are defined as follows
p

Cik=L:aisbsk (i=l, ... ,m; k=l, ... ,n). (A.2.12)


r=l

Clearly, the multiplication of matrices is not, in general, commutative.


For example, if m =I- n the product makes no sense at all. If m = n, then
ab determines a square n x n matrix whereas ba determines a p x p matrix.
Only square matrices can be commutative. For example, the unit matrix
En is commutative with any square n x n matrix

(A.2.13)

Multiplication of more than two matrices is associative. If a, b, care m x


p,p x q, q x n matrices, respectively, then (ab) c and a (bc) are equal to each
other

(ab) c = a (bc) = abc. (A.2.14)

The property of matrix multiplication is distributive over addition, i.e.

a(b+c)=ab+ac. (A.2.15)
A.2 Operations on matrices 789

Relationships (14) and (15) are proved directly by means of formula (12)
defining the operation of multiplication. This formula is applicable to the
proof of transposing the product of matrices

(ab)' = b'a'. (A.2.16)

By virtue of eq. (14) this rule is also applicable to the product of any
number of matrices

(abc)' = c'b' a'.

Let a and b be square matrices. Comparing the rule (12) with the rule
of multiplication of determinants and taking into account eq. (2) we can
write the following equalities

labl = Ibal = lallbl = la'lIbl = lallb'l = la'llb'l· (A.2.17)

Example 4. Given vectors x and y let us define a 1 x n row-matrix x' and


a n x 1 column-matrix y. Due to eq. (12) their product is a 1 x 1 matrix
which is the scalar

x'y = X1Yl + X2Y2 + ... + XnYn = y'x. (A.2.18)

Thus, the scalar product of two vectors a and b can be cast in one of
two forms

(A.2.19)

Example 5. The product xy' is the square n x n matrix

xy' = (A.2.20)

Let us recall that the dyadic product of two three-dimensional vectors a


and b is the tensor of second rank denoted by ab which has the following
table of components

(A.2.21)

This table coincides with matrix a'b, thus we can write down the following
equality

ab = ab' (A.2.22)
790 Appendix A. Elements of the theory of matrices

Postmultiplying matrix ab' by the 3 x 1 column-matrix c corresponding


to vector c we arrive, due to eq. (19), at the expression

ab'c = ab· c (A.2.23)

which determines the column-matrix corresponding to the product of vector


a with the scalar b . c. In a similar fashion, we find that

c'ab' = c· ab (A.2.24)

which is the line corresponding to the vector c . abo Therefore, we can define
the multiplication of the dyadic product ab with vector C. Post multiplying
or premultiplying the dyadic yields the vector having the direction of the
left or right vector of the dyadic, respectively.
Example 6. Let us consider the product of the skew-symmetric matrix a
corresponding to vector a by means of eq. (3) and the row b. Then we have

0 -a3 a2 b1 -a3b2 + a2 b3
ab= a3 0 -al b2 a3bl - a 1 b3
-a2 al 0 b3 -a2bl + a 1 b2
or

ab= C. (A.2.25)

Here c denotes the column-matrix with the following elements

(A.2.26)

Hence, it corresponds to the vector c which is equal to the vector product

c = a x b. (A.2.27)

Due to eqs. (7) and (16) the row-matrix corresponding to vector c is


equal to

c' = -b'a. (A.2.28)

Example 7. Let us consider the product of a m x n matrix a with a n x 1


column-matrix X. The result is the m x 1 column

Z = ax (A.2.29)

with the elements


n
Zi = Laikxk (i = 1, ... ,m). (A.2.30)
k=l
A.2 Operations on matrices 791

Premultiplying z with 1 X m row-matrix y' we obtain the scalar

"""' """'
n n
y ,z = y,ax = ~ ~ aikYiXk = x ' a,y, (A.2.31)
i=l k=l

which is a bilinear form of the variables xl, X2, ... , Xn and yl, Y2, ... , Yn. The
latter equality in eq. (31) is written, due to eq. (16), since the transpose of
the scalar is the scalar itself.
Example 8. We proceed now to the case of the square matrix and Y =
x. Then we arrive at consideration of the quadratic form of variables
Xl, X2, ... , Xn

n n
r.p (Xl, X2, ... , Xn ) = -X' ax = X' a' X = L L aikXiXk, (A.2.32)
i=l k=l

formed by means of matrix a. This quadratic form is zero if matrix a is


skew-symmetric. Indeed, byeq. (6) we have

x'ax = x'a'x = -xax' = o.


For this reason we can omit the second term on the right hand side of the
following equation

X, ax = 2x
1, (a + a ') X + 2x
1, (a - a ') x.

In other words, matrix a which produces the quadratic form can be taken
as symmetric without loss of generality.
The determinant of the quadratic matrix producing the quadratic form
is called the discriminant of quadratic form

~ = deta = lal.
Under the transformation of the quadratic form to the new variables z
introduced by means of the linear transformation

x = hz or x' = z' h'


with the quadratic matrix h, we obtain

x'ax = z'h'ahz = z'bz, (A.2.33)

where b = h' ah.


It is easy to prove that b is a symmetric matrix if a is symmetric

b' = h'a'h = h'ah = b.


792 Appendix A. Elements of the theory of matrices

Notice that, due to eqs. (17) and (2),

Ibl = Ih'llallhl = Ihl 2 lal, (A.2.34)

that is the discriminant of the form under the above transformation is


multiplied by the square of the determinant of the transformation matrix.
Matrices b and a related byeq. (33) are termed similar matrices.
5. Multiplication of block matrices. Let us consider two conform matrices
which are m x P matrix a and P x n matrix b. They can be represented as
block matrices
mi x PI mi x P2 mi x Ps
m2 x PI m2 XP2 m2 x Ps
a=
mt x PI mt x P2 mt x Ps
(A.2.35)
PI x ni PI x n2 PI x nq
P2 X ni P2 x n2 P2 x nq
b=
Ps X ni Ps X n2 Ps X nq
where

PI + ... + Ps = p, mi + ... + mt = m, ni + ... + nq = n.


This notation indicates only the orders of the submatrices. The division
is assumed to be carried out such that the submatrices in the lines of matrix
a conform to those of matrix b. Let C denote the block matrix with the
elements

C kl = (mk x PI) (PI x nl) + (mk x P2) (P2 x nl) + ... + (mk x Ps) (Ps x nl),
(A.2.36)

which are mk xnl matrices. The order of matrix C is equal to (mi + ... + mt)
x (ni + ... + nq) = m x n and coincides with the order of the product
ab = c. It is easy to prove that C = c, i.e. multiplication of two block ma-
trices implies formal multiplication of the submatrices as matrix elements.
Example 9. Let us determine the product AP of Lagrange's matrix (1.23)
and Poisson's matrix (1.28). Due to rule (36) we obtain

AP -II
-
[[aa]] ((aa)) + [[a,6ll ((,6a))
[[,6all ((aa)) + [[,6,6ll ((,6a))
[[aall ((a,6)) + [[a,6]] ((,6,6))
[[,6all (( a,6)) + [[,6,6]] ((,6,6))
I
(A.2.37)
Let us calculate the n x n submatrices of matrix AP. We have

[[aall ((aa)) + [[a,6]] ((,6a))


II~ {[as, ak] (ak' ar) + [as, ,6k] (,6k' ar)}11 (s, r = 1, ... , n)
A.2 Operations on matrices 793

Recalling the definition of Lagrange's brackets (1.20) and Poisson's brackets


(1.26) we obtain
n

k=l

Noticing that

we obtain
n

k=l
794 Appendix A. Elements of the theory of matrices

and thus

[[aall ((aa)) + [[a;3ll ((;3a)) = -En· (A.2.38)

The matrix in the right lower corner of eq. (37) differs from the above only
in permutation of a and /3, thus it is also equal to -En- We can prove that
the matrix

[[aall ((a;3)) + [[a;3ll ((;3;3)) = 0

is the null matrix, as well as the matrix which is obtained from it by means
of permutation of a and /3 in eq. (37). Hence,

(A.2.39)

Recalling that A and P are the skew-symmetric matrices we can also write
the following

A'P = Api = E 2n . (A.2.39)

A.3 Inverse of the matrix


1. The adjoint of the matrix. Let us consider lal = det a where a denotes a
square n x n matrix. It is known that the algebraic adjunct Aik of element
aik is the determinant which is obtained by removing all the elements of
the i - th row and the k - th column and multiplying by (-1) i+k.
Let us construct a n x n matrix A of algebraic adjuncts Aki

A= (A.3.1)

(please, note the order of the subindices). This matrix is referred to as the
adjoint of matrix a. Let us construct matrix aA. Its element of the r - th
row and the k - th column is equal to
n
(aA)rs = L arkAsk,
k=l

which is the sum of the products of the elements of the r - th row and the
algebraic adjuncts of the corresponding elements of the s - th row. When
r -=I- s then this sum is equal to the product of the elements of the r - th row
A.3 Inverse of the matrix 795

with the algebraic adjuncts of another row, i.e. it is equal to zero. When
r=s then this sum equals lal. Hence,

(aA)rs = brslal or aA = lalE. (A.3.2)

By analogy, we obtain

Aa = lalE. (A.3.3)

2. Inverse of the matrix. Let us assume that a is a non-singular matrix,


i.e. lal -=I- 0 and denote

A -1
(A.3.4)
~=a .

Then from eqs. (2) and (3) we obtain

(A.3.5)

The square n x n matrix a-I is referred to as the inverse of matrix a.


Clearly, only a non-singular square matrix has an inverse.
Let a and b be non-singular n x n matrices. Then by virtue of eq. (2.17)
c = ab is non-singular as well. Premultiplying both sides of this equation
by b- 1 a- 1 and post multiplying them by c 1 we obtain

or, due to eq. (2.13),

b- 1 a- 1 E = b- 1 Ebc 1 = b- 1 bc- 1 = Ec- 1 .

Thus, if ab = c then b- 1 a- 1 = c 1 .
This provides us with the formula for the inverse of a matrix

(A.3.6)

In particular
( _1)-1 =a
E -1 = E =aa (-1)-1_1
a,

and comparison with eq. (5) shows that (a- 1fl = a, i.e. the inverse of
the inverse of a matrix is the matrix itself. Notice that due to eqs. (5) and
(2.17)

(A.3.7)

Let us apply the rule of transposition to relationship (5)

E' = E = ( aa -I)' = (-1)"


a a.
796 Appendix A. Elements of the theory of matrices

Postmultiplying both sides of this equality by (a,)-I we obtain

E(a,)-I = (a-I)' a' (a') 1 = (a-I)'


or

(A.3.8)

that is, the inverse of the transpose is equal to the transpose of the inverse.
3. Solution of a system of linear equations. The system of linear equations
n
Laskxk = bs (s = 1, ... ,n) (A.3.9)
k=I

can be written in the matrix form

ax = b (A.3.1O)

by introducing a n x n matrix a and a n x 1 column-matrices x and b. If


a is a non-singular matrix then premultiplying the latter equation by a-I
we find

(A.3.11)

Recalling eq. (4) we obtain an expanded form of the solution

(A.3.12)

where the elements of the inverse e = a-I are given by


1
esk = ~Aks (s,k = 1, ... ,n). (A.3.13)

4. Adjoint quadratic form. Let

(A.3.14)

be the quadratic form obtained by means of a non-singular symmetric n x n


matrix a.
Let us introduce the new variables
orp n
ys = ax = L askxk, i.e. y = ax. (A.3.15)
s k=I

The quadratic form is then transformed as follows


1, 1,
-x ax = -y ey
2 2
A.3 Inverse of the matrix 797

where due to eqs. (2.33) and (8)

e = (a-I)' aa- l = (a-I)' E = (a-I)' = a-I.


Hence
1 1
-x'ax = _y'a-Iy (A.3.16)
22'
and expression (16) in the new variables (15) is referred to as the adjoint
quadratic form obtained by means of the inverse of matrix a.
The adjoint expression for the form rp (Xl, ... , xn) is denoted as rp' (YI,
... , Yn). Clearly

8rp' n
8 = I>skYk (s = 1, ... ,n),
Ys k=l

where the elements esk of matrix a-I are given byeq. (13).
Let us notice also the bilinear representation of the quadratic form

1, 1, 1~
'2x ax = '2x Y = '2 ~ xsYs,
k=l

where the latter equality, due to eq. (15) and (17), can be written in either
of two forms
1~ 8rp , 1 ~ 8rp'
rp='2~xs8x' rp ='2~Ys8' (A.3.19)
k=l s k=l Ys
expressing Euler's theorem on homogeneous functions in the case of quadra-
tic forms.
5. Positive definite quadmtic form. The quadratic form of n variables
1 1
L
n
rp = '2 x 'ax = '2 askXsXk
k=l

is called positive semi-definite if rp ~ 0 for all real values of the variables.


For example, the following forms

xi + x~ + X~, (Xl - X2)2 + (X2 - X3)2 + (X3 - Xt}2


satisfy this condition. A positive semi-definite form which is zero if and
only if all its variables Xl, X2, . .. , Xn are equal to zero is referred to as
the positive definite form. In the latter example, the first form is positive
definite whilst the second one is positive semi-definite because it is zero
for Xl = X2 = X3. If we consider the quadratic forms of three variables
XI,X2,X3 we conclude that the form

xi +X~
798 Appendix A. Elements of the theory of matrices

is positive semi-definite rather than positive definite since it is zero at


Xl = X2 = 0 for any X3.
The objective of the forthcoming analysis is derivation of Sylvester's
theorem that provides us with the necessary and sufficient conditions for
positive definiteness of the quadratic form. Let, first, X2 = ... = Xn = 0,
then cp = ~auxi is positive only if au > 0 and this inequality is one of the
Sylvester's criteria. Then we can write

cp (Xl, X2, ... , Xn) = 1


-au (
Xl
a 12
-X2 + + ... + -alX
n
n )2 + CP2 (X2' ... , x n ),
2 au au

where CP2 depends on variables X2, ... ,X n and does not depend on Xl. Let
us assume that

Xl + -al2
X 2 = 0, X3 = .. = Xn = 0, X2 # O.
au

Then

cP ( - :~~ X2, X2, 0... , 0) = CP2 (X2' 0, ... , 0) = ~F2X~'


and F2 > 0 and form CP2 can be represented as follows

Continuing the process we can represent form cP as a sum of products


of the positive values au = F I , F 2 , ... ,Fn with the squares of the linear
functions ~ s

~l = U12 X 2+

)
XI+ U13 X 3+ +UlnXn ,
~2 = X2+ U23 X 3+ +U2n X n,
~3 = X3+ +U3n X n, (A.3.20)
............ ............ ............
~n = Xn ,

where Uik are expressed in terms of the coefficients ask of form cpo With the
help of the new variables we can write this form as follows

(A.3.21)

Thus, the inequalities

au = FI > 0, F2 > 0, ... , Fn > 0 (A.3.22)

are the required necessary and sufficient conditions for the positive defi-
niteness of the form. These inequalities should be expressed in terms of
coefficients ask.
A.3 Inverse of the matrix 799

The linear transformation (20) can be cast in the form

~ = UX,

where u is the triangular matrix

1 U12 U13 Ul n
0 1 U23 U2n
U= 0 0 1 U3n

0 0 0 1

Its determinant is equal to unity. Introducing the diagonal matrix

o
o
F=
o o
we can put eq. (21) in the form

I.{J = i1 , F~ = "2x
1 ,
U
,
Fux

and thus

a = u'Fu. (A.3.23)

Let us designate the discriminant of form I.{J in the original variables as


bon = lal, and let bon-I, bo n - 2 , ... ,bo 1 denote the discriminants of forms
I.{Jn-l, I.{Jn-2, ... ,I.{JI which are obtained from I.{J if we consequently set Xn =
0, Xn-l = Xn = 0, ... ,X2 = X3 = ... = Xn-l = Xn = o.
It is clear that bo n - s are the principal diagonal minor determinants of
determinant bon = lal

an a12 al,n-l

bo n - 1
a21 a22 a2,n-l
, ... ,
al,n-l a n -l,2 an-l,n-l

bo 2
an
a21
a12
a22
I, bo1 = all· (A.3.24)

It remains to notice that by eqs. (23) and (2.34)

Putting ~n = Xn = 0 we obtain
800 Appendix A. Elements of the theory of matrices

For ~n = ~n-l = 0, i.e. for X n -l = Xn = 0 we have

and so on. The last equality in this chain will be

Then we find

Fl = 6. 1 , F2 = 6.
AI'
2
... ,
F n -- 6. n - 1 F - ~
,n -
U 6. n - 2 6. n - 1

and condition (22) takes the form

6. 1 > 0, 6. 2 > 0, ... , 6. n - 1 > 0, 6. n > O. (A.3.25)

This is the necessary and sufficient condition for the positive definiteness
of quadratic form <p expressed in terms of coefficients of this form.
Remark. It is easy to note from the derivation that in the case of negative
definite quadratic forms the above condition is as follows

6. 1 < 0, 6. 2 > 0, 6.3 < 0, ... ,(-It 6. n > 0, (A.3.26)

that is, the signs alternate and au = 6. 1 < O.


Let us notice in passing that the matrix of the positive or negative definite
form is non-singular since 6. n = lal =I- O.
If 6. n = lal = 0 whereas the all other 6. i > 0 for i = 1,2, ... ,n - 1,
then <p is a singular positive semi-definite form. It turns to zero at ~1 =
~2 = ... = ~n-l = 0 and any ~n = X n · For 6. n = 6. n - 1 = 0 and 6. i >
o (i = 1,2, ... ,n - 2) we have Fn = Fn - 1 = 0 and the positive semi-
definite form is the sum of n - 2 squares ~i, ...
'~~-2 multiplied by positive
coefficients F 1 , ... ,Fn - 2 , i.e. it is zero at ~1 = ~2 = ... = ~n-2 = 0 and
any ~n-l and ~n etc.
6. Orthogonal matrices. A non-singular matrix a is called orthogonal if
the inverse a-I of this matrix is equal to the transpose a' of this matrix
a-I = a' .

For the orthogonal matrix

aa' = aa- 1 = E and a'a = a- 1 a = E. (A.3.27)

By means of eqs. (2.2) and (2.17) we obtain that the square of the deter-
minant of the orthogonal matrix is equal to unity.
The product of two orthogonal matrices is an orthogonal matrix too. It
can be easily proved with the help of eqs. (27) and (6)

(A.3.27)
AA Matrix representation of the operations of vector calculus 801

A.4 Matrix representation of the operations of


vector calculus
As pointed out above, see eqs. (2.19), (2.25) and (2.27), the basic operations
of vector calculus, which are scalar, vector and dyadic products, can be
written by means of the matrix notation

a· b = a'b = b' a, (A.4.1)

c = ab = -ba, c' = -b'a = a'b, (A.4.2)

ab = ab', ba = ba'. (A.4.3)

Here c denotes the column-matrix (and correspondingly c' the row-matrix)


of projections of the vector products c = a x b and a designates the skew-
symmetric 3 x 3 matrix (2.3).
The identities

a . (b xc) = b . (c x a) = c . (a x b)
can be cast as follows

a'bc = b'ca = c'ab,

whilst the identity

a x (b x c) = b (a· c) - c (a· b) = ba· c - b· ac


is written in the form

abc = ba' c - ca'b = (ba' - Ea'b) c. (A.4.5)

As column-matrix c is arbitrary we can equate the 3 x 3 matrices in front


of c, then we arrive at the identity

ab = ba' - Ea'b (A.4.6)

and in particular

a 2 = aa' - Ea' a. (A.4.7)

Let us consider the linear functions of projections at, a2, a3 of vector a

Cl = PHal + P12 a 2 + P13 a 3, }


C2 = P2l al + P22 a2 + P23a3, (A.4.8)
C3 = P3l al + P32a2 + P33a3.
802 Appendix A. Elements of the theory of matrices

If values Cl, C2, C3 can be considered as projections of a certain vector c


on the same axes for an arbitrary vector a the following array of nine values

(A.4.9)

is said to determine tensor P, with Pik being its components along these
axes.
We can form the scalar product of vector a and tensor P

c=P·a, (A.4.lO)

vector a being called the postfactor. The matrix notation of this relation-
ship has the form

c=Pa, (A.4.11)

where P denotes the 3 x 3 matrix prescribed by the same table (9) as that
of tensor P, whereas c and a are 3 x 1 column-matrices corresponding to
vectors c and a. Transposing relation (11) yields
c' =a'P', (A.4.11)
which corresponds to multiplication of the transpose of tensor P with the
prefactor a

c=a·P'. (A.4.13)
The matrix multiplication

-a3 P21 + a2 P 31 -a3 P22 + a2 P32 -a3 P 32 + a2 P33


iiP = a3 Pll - a 1P 31 a3 P12 - a 1P 32 a3 P13 - a1P33 (A.4.14)
-a2 Pll + a1P21 -a2 P12 + a1Pn -a2 P13 + a1P23
corresponds to the tensor a x P. We can write
a x (p. b) = (a x P) . b, (A.4.15)
since both sides of this equation are equal to iiPb in matrix notation.
The matrix notation Piib corresponds to the following vector operation
(P x a) . b = P . (a x b). (A.4.15)

A.5 Differentiation of a matrix


Let us consider matrix a whose elements are functions of the variable t. By
definition of the difference of matrices we have

a (t + ~t) - a (t) = Ilaik (t + ~t) - aik (t)ll.


A.5 Differentiation of a matrix 803

1
Multiplying the result by the scalar f}.t we obtain

a (t + ~~ - a (t) = II aik (t + f}.~~ - aik (t) II.

Now we can calculate the limit for f}.t -+ 0, then the elements of the matrix
on the right hand side become equal to the derivatives of the elements of
matrix a. For this reason, it is natural to refer to it as the derivative of a
with respect to t
a=llaikll, (i=l, ... ,m;k=l, ... ,n). (A.5.1)
For example, the velocity vector v = r where r denotes the vector-radius.
Then if x and x' denote respectively the column-matrix and the row-matrix
corresponding to r, then
v= x, V
I
=
·1
x, (A.5.2)
where v and v' denote respectively the column-matrix and the row-matrix
corresponding to v.
If r depends on t both explicitly and in terms of the arguments ql, ... , qn
r=r(ql, ... ,qn;t) or X=X(ql, ... ,qn;t),
then the velocity vector is determined by the column-matrix
. ax. ax. ax
v = x = -;:;-ql
uql
+ ... + -;;-qn
uqn
+ -;:).
ut
(A.5.3)

This equality can also be written in the form


aX. ax
v = aq q + at' (A.5.4)

where q denotes the column-matrix with the elements qs, S 1,2, ... ,n
WISt aq IS t h e 3 x n matnx
h 'l ax. .

aXl aXl
aql aq2
aX2 a X2 ax
(A.5.5)
aql aq2 aq'
aX3 aX3
aql aq2
From eq. (4) it follows that
av ax ax
(A.5.6)
aq aq aq'
which is the matrix form of the following vectorial equalities
av or Or
(A.5.7)
Appendix B
Basics of tensor calculus

The analysis presented here is limited to basic knowledge. A more compre-


hensive analysis can be found, for example, in [47] and [51].

B.1 General non-orthogonal coordinates


Let us consider three non-coplanar vectors el,e2,e3 which form a vector
basis. The vectors e s are not necessarily unit base vectors, that is, the value
of e s is arbitrary. As the base vectors are non-coplanar the following value
(B.l.1)
is non-zero and can be made positive by an appropriate numbering of the
base vectors. It is equal to the volume of the parallelepiped constructed on
the base vectors.
The dual vectors are defined as follows

(B.l.2)

and form the dual vector basis. They are orthogonal to the coordinate planes
of the original vector basis
(B.l.3)
and, by virtue of eq. (1),
(B.l.4)
806 Appendix B. Basics of tensor calculus

The definitions of the original and dual bases are reversible, i.e. the basis
which is dual for the dual basis coincides with the original basis. In order
to prove this, let us calculate the vector product
1
2" (e3 x el) x (el x e2)
v
1 1
2" [eIe2 . (e3 xed - e2eI . (e3 x el)] = -el
v v
and the following expression

(B.1.5)

Then we arrive at the equalities

(B.1.6)

which completes the proof.


The following scalar products

(B.1.7)
are of crucial importance. According to eqs. (3) and (4) we have

s_{ 01
gk -
(s=k),
(s =1= k) , (B.1.8)

that is, Ilgskll is the unit matrix.


Let us prove that the symmetric matrices

are the inverse of each other. Indeed,


3 3
L e s . ek ek . em = e s . L ekek . em
k=1 k=1
3
eS'L [em x (ek X e k ) +ekem . ek]
k=1

It is easy to see that


3 1
Lek x e k = - [el x (e2 x e3) +e2 x (e3 x el) +e3 x (el x e2)] = 0,
k=1 v
(B.1.9)
B.l General non-orthogonal coordinates 807

and, hence,
3
Lgskg km = g~ or gg* = E, g* = g-l, (B.1.IO)
k=l

which is required. If follows from the above that

Gs k
ks
9 = 191' (B.1.lI)

where Gs k is the algebraic adjunct of the element gsk of the determinant


Igl of the matrix g. Alternatively, we can find that, for example,

and comparing with eq. (1) yields

v = v'T9T. (B.1.12)

In what follows we will omit the redundant summation sign provided


that the index appears twice in the summed expression, namely, once as a
subscript and secondly as a superscript. Using this notation, formula (10),
for example, takes the form
km m
gskg = gs . (B.1.13)

Using the values introduced by eq. (7) we can establish the following
relationships between the base vectors and dual vectors

(B.1.14)

where the summation signs are omitted. In order to prove this, it is sufficient
to multiply both sides of the first and second relationships by e l and el,
respectively. The result is
k
gls = gskgl = 9sl,

which completes the proof.


While using the non-orthogonal coordinates it is convenient to generalise
the Levi-Civita symbols introduced in Sec. 2.1. To this end, we introduce
two types of these symbols

E stq = e s· (e t X e)
q and Estq =e S • (e t x e q ) . (B.1.15)
808 Appendix B. Basics of tensor calculus

They are non-zero if there are no equal numbers among the indices s, t, q.
If all these indices are different and correspond to an even permutation of
1,2,3, then

= ./i9f,
Estq _ _1_
Estq - yfgT. (B.1.16)

If they correspond to an odd permutation of 1,2,3, then

Estq = -./i9f, E
stq
= -
1
yfgT. (B.1.17)

Formulae (2) and (6) are now cast in the form


es X et = cLstq e q , eS x et = cstqeq.
L (B.1.18)

B.2 Vectors using the non-orthogonal coordinates


Given a vector basis el, e2, e3, we can describe an arbitrary vector a twofold.
We can represent it in the original basis

(B.2.1)

or we can prescribe the three scalar products

as = a· e s (s = 1,2,3) . (B.2.2)

The values as and as are referred to as the contravariant and covariant


components of the vector a, respectively. Their geometrical meaning is
evident. The line segments

as lesl = as vg;;.

(no summation over s) are equal to the edges of the parallelepiped con-
structed on the base vectors. The vector a is the diagonal of the paral-
lelepiped and the line segments

are the projections of a onto the base vectors.


Using eqs. (1), (2) and (1.14) we obtain the equations relating the co-
variant and contravariant components

(B.2.3)

Now it is easy to derive the following relationships

(B.2.4)
B.2 Vectors using the non-orthogonal coordinates 809

which allows us to define the covariant and contravariant components of the


vector in the original basis as the contravariant and covariant components
of the vector in the dual basis, respectively.
In what follows we will refer to the original and the dual bases as the old
and new bases, respectively. Comparing formulae

demonstrates that the relationships between the covariant and contravari-


ant components of the vector in the old and new bases are identical to
those between the old and new basis vectors. It follows from the formulae

that the contravariant components in the old basis are related with the same
components of the new basis by the relationships of the change of basis.
This conclusion remains valid if an arbitrary triple of vectors ei, e~, e~ is
assumed as a new basis.
The difference between the covariant and contravariant components dis-
appears for an orthogonal Cartesian coordinate system with leal = 1.
The scalar product of two vectors can be set in any of three forms

(B.2.5)

In particular,

(B.2.6)

The vector product c = a x b can be written as follows

or

Therefore, the covariant and contravariant components of the vector prod-


uct are equal to

(B.2.7)

For instance,

(B.2.8)

and so on.
810 Appendix B. Basics of tensor calculus

B.3 Tensors of second rank in the non-orthogonal


coordinates
A tensor of second rank is defined by means of the nine components which
transform a vector a to another vector c. When an orthogonal Cartesian
coordinate system is used, any tensor of second rank can be presented by
means of dyadic representation (4.3.4). When a non-orthogonal coordinate
system is used the dyadic isik should be replaced by one of the following
dyadics

which gives rise to the four dyadic representations


P = P sk esek = p ske s e k = ps.kese k = p.k s
seek (B.3.1)

by means of the contravariant psk, covariant Psk and mixed p'k,p~k com-
ponents.
Using eq. (1.14) it is easy to obtain the relationships between the above
components. Post multiplying eq. (1) by em and e l we arrive at the equa-
tions

(B.3.2)

By analogy, we obtain
P1m = 9s19km P sk = 9s1 ps·m = 9km p.k
I , (B.3.3)

I
P ·m lsp plk kl p·s (B.3.4)
= 9 sm = 9km = 9 9ms k'

Pm·l = 9 sl Lms
n
= 9km pkl = 9ms9 kIps·k· (B.3.5)

These formulae explain the operations of lowering and raising of the


indices.
A tensor of second rank is symmetric if Psk = Pks' As follows from
the above formulae, the same relationship holds also for the contravariant
components psk = pks as well as for the mixed components P'k = Pi/.
Hence we can adopt the following notation P'k = Pi/ = Pi; since the
sequence of the indices is no longer needed.
An example of a symmetric tensor of second rank is the fundamental
tensor g. Its covariant 9sk, contravariant 9 sk and mixed 9'k components are
defined in terms of the base and dual vectors by means of eq. (1.7).
Given the vector a, the representation of the vector

c=P·a (B.3.6)
B.4 Curvilinear coordinates 811

can be obtained in terms of its covariant and contravariant components


with the help of the appropriate notation for P. For example, given the
covariant components of a, the covariant components of c are obtained as
follows

that is,

etc. The same strategy is applicable if P is premultiplied by a.


The tensor of second rank is referred to as skew-symmetric if psk
- pks, then Psk = - Pks and Pic = - Pi/. A skew-symmetric tensor is
given by three components. Introducing the Levi-Civita symbols, eq. (2.1),
we can assume that

(B.3.7)

Postmultiplying the skew-symmetric tensor P by a we have

c=P·a

or

(B.3.8)

which is equivalent to the following extended form

C1 = Ji9T
1 (w 2 a3 - w3 a)
2 , Cl -_ VII::Ilgl
1!l1 (w a3
2 - w3 a2 )

and so on. We can arrive at the same equation if we enter vector wand
calculate the vector product w x a. Hence, adopting notation (7) we obtain
that for the skew-symmetric tensor

P ·a=w x a. (B.3.9)

B.4 Curvilinear coordinates


The position of a point in three-dimensional space is prescribed by three
quantities ql, q2, q3, referred to as the curvilinear or generalised coordinates.
This means that the position vector r should be considered as vectorial
function of these quantities

(B.4.1)
812 Appendix B. Basics of tensor calculus

This relationship can be written in the form of three equalities

(B.4.2)

in the Cartesian coordinates.


The latter equations are assumed to be uniquely resolvable for the vari-
ables ql, q2, q3, which implies a non-zero Jacobian
OX ox ox
oql oq2 oq3
oy oy oy
J= (B.4.3)
oql oq2 oq3
oz oz oz
oql oq2 oq3

within the range of the variables ql, q2, q3. We can assume the Jacobian as
being positive by an appropriate numbering of variables qS.
For the adopted system of the curvilinear coordinates we can calculate
the following triple of the vectors

rs = -
or (s = 1,2,3). (B.4.4)
oqS

These vectors are non-coplanar as the value

(B.4.5)

is equal to the Jacobian J and, thus, is not equal to zero.


Let the vectors r sand r S be understood to be the base and dual vectors
at the point under consideration, respectively. All the previous formulae
and definitions are valid, however the quantities and the coordinate basis
change for various points in space.
Due to eq. (4) the expression for the vector dr joining two infinitesimally
close points is given by

(B.4.6)

and, thus, the square of the distance between these points is equal to

(B.4.7)

Hence, the covariant components of the fundamental tensor can be ex-


pressed as the coefficients of the quadratic form ds 2 • These determine the
metric of the chosen system of curvilinear coordinates in the vicinity of the
point under consideration, and, for this reason, the tensor g is referred to
as the metric tensor.
The formulae of differentiation of the base vectors r s and the dual vectors
r S are of crucial importance for the forthcoming analysis.
B.4 Curvilinear coordinates 813

Let us begin by calculating the following vectors

ars a 2r ark
rsk = aqk = aqkaqS = aqs = rks· (B.4.S)

They can be represented in terms of the base vectors as follows

(B.4.9)

The coefficients designated by the braces are referred to as Christoffel's


symbols of second kind. Due to eq. (S), they are symmetric with respect to
the lower indices

(B.4.lO)

If follows from eq. (9) that

(B.4.11)

and then, by virtue of eq. (1.13), we have

l } = gtl rsk· rt·


{ sk (B.4.12)

The scalar product on the right hand side is denoted as

rsk . rt = [s, k; t] = [k, s; t]. (B.4.13)

These values are referred to as Christoffel's symbols of first kind. The fol-
lowing notation

is often used for Christoffel's symbols. By virtue of eq. (13) we have

a agsk
aqtrs· rk = aqt = [s, t; k] + [k, t; s].

By means of the circular permutation we also obtain

agkt ag
aqS = [k, s; t] + [t, s; k] , aqt; = [t, k; s] + [s, k; t].

Subtracting now the first equation from the sum of the second and third
ones, and taking into account the symmetry of Christoffel's symbols with
814 Appendix B. Basics of tensor calculus

respect to the first two indices, we arrive at the formula which expresses
Christoffel's symbols of first kind in terms of the derivatives of the covariant
components of the metric tensor

[s, k.]
,t
= ~2 (88qk
gst
+ 8g
8qS
kt _ 8 9Sk )
8qt . (B.4.14)

With the help of eqs. (12) and (13) we have

{slk} = gZt [s, k; t]. (B.4.15)

The sequence of eq.(15) is the following equalities

[s, k; t] = gtm {:}. (B.4.16)

A simple way to obtain the derivatives

(B.4.17)

is as follows. Due to eq. (9) we have

8 t =O=-rz·r
-gz
8qS
8
8qS
t
=rz·rt
S
+ sl{m} rm·r t

or

(B.4.18)

This means that the values on the right hand side can be formally treated
as the covariant components of the vectors r~. Thus

(B.4.19)

Let us notice that Christoffel's symbols are zero if and only if gtm are
constant values, i.e. when the base vectors r S retain their values and direc-
tions.

B.5 Covariant differentiation


In mechanics and mathematical physics, the invariant quantities are of in-
terest. They do not depend on the coordinate basis and are determined
by the properties of the object under consideration. The invariants can be
B.5 Covariant differentiation 815

scalars (for example, energy, work, mass, temperature etc.), vectors (ve-
locity, acceleration, force) and tensors (inertia tensor, strain tensor, stress
tensor), as well as functions of the above invariants, for example, dyadic,
scalar and vector products of vectors, tensors and so on.
In order to carry out calculations with the vectorial and tensorial quanti-
ties we need to introduce the covariant basis and the components of vectors
and tensors (contravariant, covariant or mixed) with respect to this basis.
Any change of the invariant in time and space reflects the property of this
invariant. The situation is different if the components are considered. Their
change is also caused by the change in the values and directions of the ba-
sis vectors. For example, let as be independent of the coordinates, that is
the derivatives with respect to the coordinates are equal to zero. However
it would be a grave error to think that the vector a does not change in
space. The inverse statement is also valid, namely, the components as do
not retain constant values for a constant vector a. The aim of the forth-
coming analysis is to introduce such characteristics of the components of
the tensors and vectors which reflect changes both in these quantities and
the vector basis. This aim is achieved by introducing operations of the
covariant differentiation.
Let us consider the derivative of vector a with respect to one of the
variables qS. To begin with, we describe the vector by its contravariant
components, then

8a 8 k
8 qs = 8qS ark·

It is clear that the formal rules of differentiation of a sum, a product etc.


remain valid when an invariant, say a vector product or a dyadic product,
is differentiated. Taking into account formulae (4.9) we have

k m
88a
qs = rk 8a
8qS + a k{m}
sk rm = rm (8a
8 qs + {m}
sk ak) . (B.5.1)

Using formulae (4.19) we also obtain that

-8a -_ r k -
8qS
- - ak {
8ak
8qS
k}r
sm
m -
_ r m (8a
--
8qS
m
- {k}
sm
)
ak. (B.5.2)

hence,

(B.5.3)

where the expressions

m
V'sa
m
= 8a
8 q s + {m}
sk a k , (B.5.4)
816 Appendix B. Basics of tensor calculus

are called the covariant derivatives of the contravariant and covariant com-
ponents of the vector a respectively.
This calculation can be easily generalised to tensors of any rank. Let us
restrict our consideration to the tensors of second rank. Using the dyadic
representation of the tensor, we obtain

We can do the same if the tensor is described by its covariant or mixed


components. Then we obtain

8~s P = rkrt "V spkt = rkrt"V sPkt = rkrt"V sp.~, (B.5.5)

where the covariant derivatives of the quantities pkt, Pkt, P.1 are given by

"Vspkt = 8pkt +{ k }pmt +{ t }pkm,


8qs sm sm

"VsPkt = 8Pkt
8qS - {m}
sk Pmt - {m}
st Pkm, (B.5.6)

"V p k
s·t
= 8P.1
8qS
+ { sm·
k }pm _
t
{m}pkm .
st·

These formulae enable the covariant derivatives of the components akbt of


a dyadic to be calculated. It is easy to see that the rule

remains valid under the covariant differentiation. This rule is also valid
when the covariant differentiation of the product of the tensor components
is carried out. For example,

The covariant derivative of a scalar is given by


8<p
"V S<P = 8qs·

Due to eq. (6), in the case of the scalar product of two vectors we have

as the second and third terms cancel out.


B.5 Covariant differentiation 817

The Ricci theorem plays an important role in tensor calculus. It states


that the covariant derivatives of the covariant, contravariant and mixed
components of the metric tensor are equal to zero. Indeed, applying the
second formula in eq. (6) to quantities gkt we obtain, due to eqs. (4.16) and
(4.14), that

V'sgkt = Ogkt
oqS -
{m}
8k gmt -
{m}
8t gkm = ogkt
oqS - [8, k; t] - [8, t; k] = O.
(B.5.7)
The mixed components are either equal to zero or unity, thus,

og~ =0
oqS .

Calculating the covariant derivative by means of eq. (6) we obtain

It remains to consider the contravariant components. Using relationship


(1.13) we obtain, by virtue of eq. (7), that

V' sg~ = 0 = V'sgkmg mt = gkm V' sgmt .


Given 8 and t, we obtain a homogeneous system of linear equations with
the non-zero determinant Igl, hence
V'sgmt = o.
Summarising we see that
V'sgkt = 0, V'sgkt = 0, V' sg~ = O. (B.5.8)
It follows that the components of the metric tensor under the covariant
differentiation exhibit the behaviour of constant values, that is, they can
be taken out of the operation V's or entered into operation V's. We can see
from formulae (5) that

Let us consider now the tensor


s oa 1 oa 2 oa 3 oa (B.5.9)
r~=r !::Il+ r !::I2+ r !::I 3'
uqs uq uq uq
which is a sum of three dyadics. Formulae (3) yield the dyadic representa-
tion of the tensor

(B.5.1O)
818 Appendix B. Basics of tensor calculus

The values \7 sam and \7 sam are the coefficients of the dyadic representa-
tion of tensor (9). Thus, they can be considered as the c:omponents of the
tensor, namely, \7 sam are the mixed components (covariant and contravari-
ant with respect to sand m respectively), whereas \7 sam are the covariant
components with respect to both indices.
The partial derivatives of both a i and ai with respect to qS are not the
tensor components.

B.6 Examples of non-orthogonal curvilinear


coordinates
In order to illustrate the calculation let us consider the following exam-
ple. A plate, whose plain is always parallel to the plane Oxy, moves along
axis Oz and rotates about it, with the rotation angle being given cp = TZ
(T is constant). Let two orthogonal axes OXI and Ox 2 lie in the plane
of the plate, then any point in space is typified by the three coordinates
xl, x 2 , x 3 = Z which can be understood as ql, q2, q3. Equation (4.2) ex-
pressing the Cartesian coordinates x, y, Z in the fixed axes in terms of the
generalised coordinates takes the form
x = xl cos TX 3 - x2 sin TX 3 , Y = Xl sin TX 3 + x 2 cos TX 3 , Z = x 3.
(B.6.1)
Thus, the projections of the base vectors of the covariant basis on these
axes are given by
rl: cos TX 3 sin TX 3 0
r2: - sin TX 3 cos TX 3 0
r3: -TY TX 1
The covariant components of the metric tensor are equal to
911 = 1, 912 = 0,
922 = 1, (B.6.2)

Now we have
1 o
o 1 TXl
=1 (B.6.3)
-TX 2 Txl 1+T2[(xl)2+(x2)2]

and
9 12 = -TX l X 2 ,
9 22 = 1 + T2 (xl) 2 , (B.6.4)
B.7 Formulae of the theory of surfaces 819

All brackets which do not contain index 3 are equal to zero. The non-zero
components are listed below

[3 , 3·, 1] = _XlT2 , [3,3; 2] = _x2T2 }


[2,3; 1] = [3,2; 1] = -T, [1,3; 2] = [3,1; 2] = T, (B.6.5)
[1, 3·3]
, = [3 , 1·3]
, = XlT2 , [2,3; 3] = [3,2; 3] = _X2T2.

It remains to construct expressions for Christoffel's symbols of second


kind by means of eq. (4.15). The result is

(B.6.6)

whilst the other symbols are zero. The calculation can be simplified by
direct determination of the derivatives of the base vectors. This yields

rl h cos Tx3 + i2 sin TX 3 , r2 = -il sin Tx3 + i2 cos TX 3 ,


i1 rl cos Tx3 - r2 sin TX 3 i2 = rl sin Tx3 + r2 cos TX 3 ,
r3 T(-ily+i2X)=T(xlr2-x2rl). (2.6.7)

For this reason

r13 = T (-il sin Tx3 + i2 cos TX 3 ) = Tr2, }


r33 = T (xlr23 - x2r13) = -T 2 (xlrl + x2r2) .

(B.6.8)

Now, using the formulae in eq. (4.9) we arrive at the same expressions
(6) for Christoffel's symbols of second kind.

B.7 Formulae of the theory of surfaces


Denoting the radius-vector and the Gaussian coordinates of a point on the
surface by p and ql, q2, respectively, we can cast the vectorial equation of
the surface in the form

(B.7.1)

We can now repeat the above derivations, the Greek indices being equal to
1 and 2.
First, the base vectors on the surface

(B.7.2)
820 Appendix B. Basics of tensor calculus

are determined. Their directions coincide with the tangents to the coordi-
nate lines of the surface. Then we can determine an infinitesimally small
displacement on the surface and the square of its value as follows

(B.7.3)

where

(B.7.4)

are the covariant components of the metric tensor of the surface. The con-
travariant components are calculated by analogy with eq. (1.11) and are
given by
12 a12 22 au (B.7.5)
a =-~' a =~'

where JaJ is the determinant

JaJ = I au
a12
(B.7.6)

and the following equality holds

(B.7.7)

The dual vectors

(B.7.8)

are also introduced into consideration. In a similar manner to the base


vectors they are the surface vectors, i.e. they lie in the tangent plane to the
surface. The equality

(B.7.9)

holds, that is pI and p2 are perpendicular to PI and P2. Let us notice that

(B.7.1O)

i.e. the element of the surface area is equal to da = Jjafdq 1 dq2.


Any vector c on the surface (i.e. the vector lying in the plane tangent
to the surface) can be prescribed by the contravariant COl and covariant COl
components

(B.7.11)
B.7 Formulae of the theory of surfaces 821

where

ca = C . P a' ca = C . pC< . (B.7.12)

In order to define the operation of differentiation of vectors on the surface,


we need the vector of the normal to the surface since the derivative of the
surface vector is a vector that does not lie on the surface.
The unit vector of the normal to the surface is defined as follows
PI x P2 1
m (B.7.13)
IPI X P2 I
= JT::iPI
vial
X P2'

hence,

Pa ' m = O. (B.7.14)

taking derivative with respect to q{3 we obtain

(B. 7.15)

The quadratic form of differentials dqa of the Gaussian coordinates

(B.7.16)
are referred to as the first and second quadratic forms of the surface re-
spectively. It will be shown below that both quadratic forms have invariant
geometrical meaning. The coefficients ba {3 of the second quadratic form are
the covariant components of the tensor, the mixed components are

bJ = a'Y a ba {3 = -a'Y a pa . m{3 = _p'Y . m{3, (B.7.17)


and the contravariant components are given by

bti {3 = a 15a a{3'Yb{h' (B.7.18)

We proceed now to obtaining formulae for the second derivatives P a {3 of


the radius-vector p. As indicated above, the vector Pa {3 is not a surface
vector and, thus, its representations should contain a component along the
normal to the surface. Assuming

Pa{3 = {a:~ }P'Y + ba{3m (B.7.19)

and taking into account eqs. (14) and (15) we find that ba {3 are the intro-
duced coefficients of the second quadratic form. Then, due to eqs. (14) and
(4) we obtain

(B.7.20)
822 Appendix B. Basics of tensor calculus

Repeating the derivation carried out in Sec. B.4 for the three-dimensional
case we obtain
1 (8aao
Pa{3 . Po = [a,,6; 8] ="2 8q{3 + 8a{3o
8 qa -
8aa{3 )
8 qo . (B.7.21)

It follows from eqs. (20) and (7) that

{a1',8} = al'O [a,,8; 8]. (B.7.22)

The values

[a,,8; 8] and {1'a,6}


are Christoffel's symbols of first and second kind for the surface under
consideration. They are calculated in terms of the coefficients of the first
quadratic form. Conversely, the values of ba {3 can not be expressed in terms
of these coefficients.
It is necessary to add the relationships
rna ·rn=O, (B.7.23)
which indicates that the derivative of the unit vector is orthogonal to this
vector. For this reason, rna, being a surface vector, can be represented in
terms of the vectors p{3 or P{3. Accounting for eqs. (12), (15) and (17) we
have
(B.7.24)

Let us construct the formulae for differentiation of the dual vectors, i.e.
a 8pa
P{3 = 8q{3·
Repeating the derivation resulting in formula (4.18) we find that

p~. Pl' = -{;1'}. (B.7.25)

Then we have

p3 . m = (8~{3aal' Pl') . rn = aal'bl'{3 = b~, (B.7.26)

as the other components of the vector obtained by differentiating aal' Pl'


are orthogonal to m. The consequence of eqs. (25) and (26) is the required
relationship
B.8 Curvature of lines on the surface 823

B.8 Curvature of lines on the surface


The line on the surface is given by the equations

(B.8.1)

expressing the Gaussian coordinates of the surface as functions of a single


parameter t. The vector equation for the line on the surface is thus cast in
the form of eq. (7.1) under the assumption that ql and q2 are given by eq.
(1 ).
In what follows, the arc length CJ of the line in question is taken as an
independent variable. The unit vector 7" of the tangent is equal to
dp udqa
7"=-=p - . (B.8.2)
dCJ dCJ
Denoting the curvature of the line and the unit vector of the principal
normal by k and n, respectively, we obtain with the help of Frenet's first
formula
d 2P d7" d 2qa dqa dq(3
dCJ2 = kn = dCJ = Pa dCJ2 + Pa(3 dCJ dCJ'
Replacing here Pa (3 by means of eq. (7.19) we have

(B.8.3)

The vector of curvature kn has a component along the normal to the surface

- dqa dq(3 ba(3dq a dq(3


mk = mkn . m = mba(3 -d -d = m d (3' (B.8.4)
CJ CJ aa(3dqa q

which is referred to as the vector of the normal curvature, and a component


in the tangential plane.
The normal section of the surface is a planar curve obtained from in-
tersection of the surface by the plane spanned by the surface normal m
and the surface tangent 7". Due to Meusnier's theorem, the curvature of
the normal section of the surface is given by

k= kn . m = k cos (). (B.8.5)

The second component in the expression for the curvature vector

k * =kn-km=p
_ (d 2qa
--+ {a}d -q(3 dq
- 'Y) (B.8.6)
a dCJ 2 (3"( dCJ dCJ

is called the vector of the geodesic curvature. The values in parentheses


are the contravariant components of this vector. They are zero along the
824 Appendix B. Basics of tensor calculus

geodesic lines on the surface. The differential equations for the geodesic
lines have the form

d 2 qO +{ a }dqf3 dq'Y = 0 (a = 1,2). (B.8.7)


drY 2 (3"( drY drY

A single geodesic line passes through any surface point in any direction.
It follows from the theorem on the existence of the solution of the system
of two differential equations (7) for given initial values of functions qO and
' d . t'
thelr dqo
enva lves drY'
It also follows from the above that the geodesic lines can be defined as
the surface curves whose principal normals have the direction of the normal
m to the surface. Then n = ±m and k = ±k, that is, k* = O.
The geodesic curvature of the surface line is the value of the vector of
the geodesic curvature, i.e.

(B.8.8)

It is equal to zero along the geodesic lines of the surface.


The contravariant components of vector T are used in the above deriva-
tion. The covariant components can be represented in the form

(B.8.9)

where <I> denotes the following quadratic form

1 dqo d q f3
<I> = '2aof3 drY drY' (B.8.1O)

Using formulae (7.27) we obtain

kn =

or
B.9 Covariant derivative of a vector on the surface 825

Recalling the relationship

and noticing that

~ (8af3ti + 8aati _ 8af3a ) d qf3 dqti


2 8qa 8 qf3 8qti dcr dcr
~ 8af3ti dqf3 dqti 8<1>
2 8qa dcr dcr 8qa '

we obtain that

(B.8.11)

Expressions in the brackets are the covariant components of the vector of


the geodesic curvature and are zero along the geodesic lines. Hence, another
form of the differential equations of the geodesic lines is available

!:...- 8<1> _ 8<1> =0 ( a=I,2.


) (B.8.12)
dcr 8 (dd~ ) 8qa

We notice also that vector T and the vector of the geodesic curvature are
orthogonal since

T . k* = T . (kn - kmm . n) = O. (B.8.13)

B.9 Covariant derivative of a vector on the surface


Let us construct the expression for increment de in the vector e prescribed
on the surface by its contravariant components en. Applying the formula
of differentiation (7.19) we obtain

Be f3 _ f3 8 "
de = 8 qf3 dq - dq 8 qf3 c Pa

d qf3 p" (8c"


8 qf3 a } c'"Y) + d qf3 c"b "f3m.
+ { ryj3 (B.9.1)

The term

(B.9.2)
826 Appendix B. Basics of tensor calculus

is the component of de along the normal to the surface, whereas the term

(B.9.3)

is a surface vector. Here

(B.9.4)

denotes the covariant derivative of the contravariant quantities ca.


Along a line L on the surface we have

dq{3
dq{3 = -da
da '
and the expression for d* e takes the form

d * c -- daPa (ac
a
aq{3
q(3)
+ { 'Ya/3 } c'Y dda (a + {a}
-_ Pa dc 'Y/3 c'Y dq (3) . (E.9.5 )
The vector is said to have translated parallel along the line L if

d*e = 0 i.e. \7 (3c a = o. (B.9.6)

In particular, when e is the vector of the tangent to the line L, then, due
to eq. (8.6),

(B.9.7)

With this in view, let us consider the change in the scalar product e· r
along L. Taking into account that e and r are the surface vectors we obtain

e .m = 0, r .m = 0

and, therefore,

d(e· r) = e· dr + r· dc = c· d*r + r· d*e. (E.9.8)

It follows that under a parallel translation of e along the line L we have

d(e· r) = e· k*da, (B.9.9)

and, specifically, when L is a geodesic line, then

d(e·r)=O. (B.9.1O)

If c is a vector of constant length, then its angle with the geodesic line
is constant under the parallel translation. For example, under a parallel
B.9 Covariant derivative of a vector on the surface 827

translation of a vector along a straight line in the plane, the angle between
this vector and the straight line remains unaltered.
Describing vector c by its covariant components we have

d* c -d
- qf3 p ar7
V f3c a -d
- qf3 p a (ac
a qaf3 - {j3a
' } c'¥ ) , (B.9.11)

where the values in the parentheses are the covariant derivative of the co-
variant quantities Ca. Repeating the reasoning of the above transformation
we obtain

d* c=ad pa (dc a _ { I } dqf3 ) (B.9.12)


da j3a da ~ ,

that is, under the parallel translation of vector c along the line L on the
surface

(B.9.13)

Knowledge of the arc length on the surface determines the first quadratic
form of the surface and in turn the inner geometry of the surface. This ge-
ometry is defined by the components of the metric tensor and all parameters
derived from it, i.e. the area element, Christoffel's symbols of first and sec-
ond kind and the geodesic curvature of the line on the surface. The problem
of determining the geodesic lines, operations of the covariant differentiation
and the parallel translation of the vector also belong to the inner geometry.
All the above quantities remain unaltered under the bending of the surface
which is not accompanied by a change in the length of the surface lines.
The normal curvature changes under bending and this indicates the fact
that the coefficients of the second quadratic form can not principally be
calculated in terms only of the metric tensor. Their definition is associated
with introducing the vector of the normal to the surface.
Let the surface under consideration correspond to a value q3 = q& of the
spatial system of orthogonal curvilinear coordinates ql, q2, q3

Then

and

m (B.9.14)
828 Appendix B. Basics of tensor calculus

Here the relationship (1.2) is used. Now we have

381 1 8r~
m/3 = ro 8 /3 1:33
q y90
+ y9()
1:33 8 /3'
q
(B.9.I5)

Noticing that

and accounting for

we arrive at the equality

(B.9.I6)

Here

(B.9.I7)

and, in order to calculate [a, (3; 3J o' we need the derivative of 901./3 with
respect to q3. Thus, determination of the coefficients bOl./3 implies knowledge
of the components of the metric tensor 901.3 at q3 = q5 and the components
of 901./3 up to terms of the first order in q3 - q5

901./3 = aOl./3 + ( 8901./3)


8 q3 0 (3
q - qo3) + ... (B.9.I8)

We have to "leave" the surface as it is not possible to determine the second


quadratic form while "staying" on it.

B.lO Orthogonal curvilinear coordinates


When the base vectors rl, r2, r3 are orthogonal, the non-diagonal compo-
nents 9sk of the metric tensor are zero. Denoting the diagonal components
9ss by h~ we have

(s =F k) ,
(B.IO.I)
(s = k),
where hs are referred to as Lame's coefficients. Then we have

(B.IO.2)
B.IO Orthogonal curvilinear coordinates 829

and thus
ss 1 s rs
g = h2 ' r = h2 • (B.lO.3)
s s

The expression for the vector a can be cast in either of the following forms

(B.lO.4)

The physical components a(s) are the coefficients of the unit base vectors
is
• rs h s
(B.lO.5)
Is = hs = sr,

Therefore

(B.lO.6)

with no summation over s.


The square of the element of the arc length has the form

(B.lO.7)

that is, Lame's coefficient is the factor of the arc length of the corresponding
coordinate line.
The non-zero Christoffel's symbols of first and second kind are listed
below

[ ) ohs s = { s } = olnhs
s, s; k = -hs oqk ' ks oqk'
sk
k hs ohs
[s, k; s) = [k, s; s) = hs ~~: ' ss -h~ oqk'
(B.lO.8)

[ ) ohs s olnhs
s,s;s = hS8'
qS oqS
ss
As an example, let us consider the spherical coordinates R, {), <p, where
R denotes the radius of a sphere on which the point lies, {) is the angle
along the meridian from the north pole (0 :S {) :S 71") and <p is the azimuth
of the meridian from the plane Ozx to the plane Oyz (0 :S <p :S 271") .
The Lame's coefficients are equal to

hR = 1, h?9 = R, h<p = Rsin {), (B.lO.9)

as the lengths of the arc element of the radius, meridian and the parallel
circle are dR, Rd{), Rsin {)d<p, respectively. The base vectors are given by

(B.lO.lO)
830 Appendix B. Basics of tensor calculus

where the unit base vectors i R , i'!9, icp have positive directions of the radius,
the tangent to the meridian and the tangent to the parallel circle respec-
tively. Now we have

There is no need to repeat the derivations carried out for the orthogonal
Gaussian coordinates on the surface. The unit base vectors ~l Pl' ~2 P2' m
form an orthogonal trihedron, so that

hl
Pl = h 2 P2 x m,

It follows that

Let us consider the three vectors Pl' m and m* = m + m1dql. The latter
is the vector of the normal to the surface at the point (ql + dql, q2) which
is infinitesimally close to the point (ql, q2) under consideration. The vector
product

which is proportional to b12 , is the volume of the parallelepiped constructed


on the vectors Pl,m,m*. It is zero when these vectors are coplanar, which
takes place if and only if the vectors m and m* are parallel or intersect.
This property defines the lines of the surface curvature. For instance, these
are the meridians and the parallel circles of the surface of revolution. As
follows from the above, b12 = 0 if the curvature lines are assumed to be
the coordinates line, however, it is true not for any mesh of the orthogonal
lines on the surface.
B.IO Orthogonal curvilinear coordinates 831

Turning to formulae (9.17) and assuming that the surface in question


corresponds to the value q3 = qg of the orthogonal system, we obtain

3
go' = 0,
j;i33 _ 1
9o - (h )0' [1,2;3] = 0, [a, a; 3] = -he> (~~~ ) 0 '
3

with no summation over s. Hence,

(B.10.12)

The latter equality shows that the surfaces of the orthogonal system in-
tersect on their curvature lines. Assuming ql = 13, q2 = rp, q3 = R for the
spherical surface of the radius R o, we obtain

b{}{} =- R o, b<p<p =- Ro sin 2 13

and the normal curvature, due to eq. (8.4), is given by

k_ b{}{} (d13)2 + b<p<p (drp )2 1


- h~ (d13)2 + h~ (drp)2 Ro

In accordance with eq. (11), there are only two non-zero Christoffel's sym-
bols

and the covariant components of the vector of the geodesical curvature are
as follows

For example, along the meridian

drp
da = Rod13, da = 0, i.e. k*{) = k*<P = 0,

i.e. the meridian is a geodesic line. For a parallel circle

da = Ro sin 13drp, k*{}


1
= - - cot 13
R5 ' k*<P = °
and, due to eqs. (8.8) and (8.3)

k* = cot 13
Ro'
k = Jk + 2 k*2 = 1
Ro sin 13·
832 Appendix B. Basics of tensor calculus

As a second example, let us consider the surface of a right circular cone


Then, by virtue of eq. (12), we have
{} = {}o.

and the normal curvature of the line on the cone is


k = _ cot {}o _ _ _ _ _1_ _ _---,,-
R 1 + __1....,,-_ (dd~)2.
R2 sin 2 {}o 't'

The non-zero Christoffel's symbols are

The contravariant components of the vector of the geodesic curvature are

k*'P = ~c.p + .3.. dc.p dR _ d 2R


k *iJ - - - - R.
sm 2 {}0 ( dc.p )
-
2
dO' 2 RdO' dO" d2O' dO'

and the geodesic curvature of the parallel circle (dO' = R sin {}odc.p) is R- 1 •
Let us also study the case of a surface of revolution. The position of a
point on the surface is given by the arc ql = 8 along the meridian and the
azimuth q2 = c.p which is the angle between the meridian plane and the
assumed plane q5
= O. Lame's coefficients are given by

hl=l, h2=r(8),

where r (8) denotes the distance of the point from the axis of revolution.
The non-zero Christoffel's symbols are

and the equations for the geodesic lines on the surface of revolution take
the form

d2c.p r' dc.p d8


da 2 +2-- -d =0,
r da a

where a is the arc along the geodesic line. The first two integrals are as
follows

and the solution reduces to the two quadratures.


B.ll Finite-dimensional Euclidean space 833

B.II Finite-dimensional Euclidean space


The subjects of analytical geometry are sets consisting of number, pairs of
numbers and triples of numbers. The algebraic results obtained are inter-
preted geometrically.
A natural generalisation leads to consideration of a manifold whose ele-
ments are m numbers Xl, ... ,X m . Let us agree to understand these num-
bers as the coordinates of a point in a Cartesian orthogonal system of axes,
then the set (Xl, ... ,xm ) corresponds to a space of the dimension m. If
the distance M N between two points of this space M (Xl, ... ,xm ) and
N (Yl,'" ,Ym) is given by
-M N V(Xl - yd
= 2 + ... + (Xm - Ym) 2 , (B.Il.l)
then this space is referred to as the Euclidean space Em.
A detailed presentation of the definitions and conclusions which enable
generalisation of the concepts of the line segment, vector, angle, area, vol-
ume etc. is out of place in the present book. It is sufficient to mention that
the customary geometrical language of E3 is valid in Em. For example, the
coordinate transformation with the same origin 0 is carried out by means
of the following formula
m

x' = ax or x~ = L askXk, (B.ll.2)


k=l

where a is an orthogonal m x m matrix describing the transformation of


one coordinate system to another (rotated) system. Here x' and X denote
the row matrices of the new and old coordinates of the point M. The
element ast of the rotation matrix is called the direction cosine of the
angle between the axes Ox~ and OXi, whereas the values ast for a taken
s determine the projections of the unit base vector i~ of the axis Ox~ on
the axes OXl, ... ,Oxm of the old system. Geometrically, the orthogonality
of the unit base vectors i~ corresponds to the following property of the
orthogonal matrices
m

aa' = E or L astakt = 8sk .


t=l

The set as, transformed under the change from the old basis to the new
one by means of formulae (2), determines a vector a in Em. The values
as are called its projections. In particular, Xs are the projections of the
radius-vector r. When Xt = 8ts (for a fixed s), then the radius-vector is the
unit base vector is of the axis OX s . Any vector can be expressed in terms
of the unit base vectors

(B.Il.3)
834 Appendix B. Basics of tensor calculus

The scalar product of two vectors is defined as follows


m
a· b = Lasbs. (B.Il.4)
s=1
The concept of the tensor of second rank is introduced in Em by analogy
with E 3 . It is a set of m 2 values Pst which is transformed due to the rule
m
bs = LPstat, (B.Il.5)
t=1
when at and bs are the projections of the vectors a and b. The tensor can
be prescribed in terms of dyadic products
m m

P = L L Pstisit . (B.Il.6)
s=1t=1
An example of the tensor is the dyadic product of the vectors
m m

ab = L L asbtisit . (B.Il.7)
s=1t=1
A curve in Em is given by the parametric equations

xs = Xs (t) and r = r (t) . (B.Il.8)

For instance, the arc IJ along the curve can be assumed as the parameter,
then the differential of the arc, due to eq. (1), is determined as follows

m
dIJ = L (dx s)2 = (B.Il.9)
s=1

The vector
dr
T= - with the projections
dIJ
determines the unit vector of the tangent to the curve. Its derivative defines
the vector referred to as curvature vector kn, which has the direction of n

kn = dT = d2 r (B.Il.l1)
dIJ dIJ 2
where n denotes the unit vector of the principal normal to the curve. The
generalisation of the vector products and other concepts is more difficult.
B.11 Finite-dimensional Euclidean space 835

It is clear that the non-orthogonal coordinate systems with the base


vectors e s can also be introduced. They determine the tensor with the
covariant components
gsk=es·e s (s,k=l, ... ,m). (B.11.12)

Its contravariant components gsk are determined by relationships (1.11) as


elements of the inverse for matrix Ilgsk II. By virtue of the properties of the
inverse for a matrix, the mixed components
(B.11.13)
are given byeq. (1.8). Then, eq. (1.14) yields the vectors of the dual basis

so that we obtain, with the help of eqs. (1.12) and (1.13), that
e S • et = gt, e S . e t = gSllqglq = gSlgj = gst. (B.11.14)
The vector a can be prescribed by means of one of the following presenta-
tions
(B.11.15)
i.e. by means of contravariant as or covariant as components. In contrast to
Sees. B.1-B.3 the indices, the dummy ones included, take values 1, ... , m
rather than 1,2,3.
The curvilinear coordinates ql, ... , qm can also be introduced in Em. The
content of Sees. B.4 and B.5, the definitions of Christoffel's symbols and
the covariant differentiation included, can be adopted here. It is essential
to remember that the square of the arc differential
(B.11.16)
can always be transformed to the sum of squares
(ds)2 = (dxd 2 + ... + (dXm)2 . (B.11.17)
To this end, it is sufficient to return to the Cartesian coordinates Xs by
means of the equations which are the inverse to the following ones
(B.11.18)
This inversion is always possible when the determinant
OXI OXI
oql oqm
M= oXm oXm
(B.11.19)

oql oqm
is non-zero.
836 Appendix B. Basics of tensor calculus

B.12 Riemannian space of dimension n


The relationship

XS=XS(ql, ... ,qn) (s=l, ... ,n, ... ,m) (B.12.1)

or in vectorial form

(B.12.2)

where n < m, is said to determine the Riemannian space Rn in the Eu-


clidean space Em. For example, a surface in E3 is considered as space R 2 .
Let the Greek letters denote the numbers 1, ... , n and the indices whose
values are greater than n be designated by n + t, n + q etc., the Latin letters
t, q etc. taking values of 1, ... , m - n.
The relationship

dr = or d 0:
oqO: q
= op d 0:
oqO: q (B.12.3)

separates the set belonging to Rn from the set of infinitesimally small


vectors at the point M (Xl, ... , xm). The vectors

op
Po: = oqO: (B.13.4)

form a coordinate basis of the space Rn in the vicinity of point M. As Rn


belongs to Em, the square of the arc element is given by

(d S ) 2 _
-
()2
dXl + ... + (d)2
Xm
_ Op 0: Op (3 _ 0: (3
- oqO: dq . oq(3 dq - ao:(3dq dq .
(B.12.5)

Let us note that our consideration is limited to the spaces with a positive
quadratic form (dS)2. The special theory of relativity also deals with the
generalised Euclidean coordinates with the quadratic form f=
0:=1
Co: (dxo:) 2 ,
a few of the constants Co: being negative. In the latter equation

(B.12.6)

denotes the covariant components of the metric tensor. Its contravariant


components ao:(3 are determined as elements of the inverse to the matrix
II ao:(3ll· The inverse exists as the quadratic form (5) of the differentials dqO:
is positive definite. Given ao:(3, we can construct the dual basis

(B.12.7)
B.12 Riemannian space of dimension n 837

at point M. A vector e is said to belong to the Riemannian space R n , if it


can be expressed in terms of the vectors Po or the dual vectors po

(B.12.8)

The essential difference between the equations (11.16) and (5) for the
square of the arc differential in Em and Rn is that in the first case there
exists a transformation of q8 to m variables Xl, ... ,X m (the Cartesian
coordinates) in which this quadratic form can be expressed as a sum of
squares, see eq. (11.17). In general, form (5) can not be expressed as a sum
of the same number n of the new variables in the whole space Rn. This is
possible only in the vicinity of a fixed point qg since we can introduce (not
uniquely) such linear functions po of the differential dq f3

t3dqf3 = dpo (B.12.9)

that the expression for (ds)2 takes the form

(B.12.1O)

System (9) is not integrable in the general case as po are differentials of


the quasi-coordinates. Hence, only a local expression for (dS)2 in form (10)
is possible since relationships (9) are integrable for constant values of t3, Le.
for fixed values of ql, ... ,qn. In the infinitesimally small neighbourhood of
this point, the geometry of Rn is Euclidean. Complication arises when we
compare the quantities at two different points of Rn (Le. Em). As illustrated
in Secs. B.7-B.9 by an example of the manifold R2 in E 3 , it is necessary to
differentiate the internal properties of R2 due to the given tensor a o f3 from
the relationship between Rn in Em.
The operation of the covariant differentiation is determined in Rn as
follows

(B.12.11)

where index a indicates that Christoffel's symbols are calculated in the


metric of Rn

{a}
(3"( a
=
a
00 [(3 . 8]
, "( ,
= ~
2a
00 [aaf3O
aq'Y +
aa'Y o _ aaf3'Y]
a qf3 aqo· (B.12.12)

Using eqs. (11), (9.5) and (9.12) we can construct the expressions for the in-
crement d*e of vector e in the space Rn when it moves to an infinitesimally
close point of the line qO = qO (a). However it is necessary to remember
that the increment de of this vector in Em differs from d*e in the compo-
nents which are normal to the base vectors Po of the space Rn at the point
838 Appendix B. Basics of tensor calculus

under consideration. This point is considered in the next section at greater


length.
The definitions of the geodesic curvature, geodesic line and parallel trans-
lation of vectors introduced in Sees. B.8 and B.9 for R2 remain valid in the
space Rn as well.
If one of the base vectors, say P6, is assumed to replace the vector c in
eq. (9.5), then the covariant components ca. should be replaced by

(aio 8),
(a = 8).

This yields the relationship

(B.12.13)

The introduced vectors

(B.12.14)

can be referred to as the covariant derivatives of the base vectors P6 in the


space Rn. Due to the symmetry of Christoffel's symbols we have

"V (3P6 = "V6P(3· (B.12.15)

Relationships (14) replace equations (4.9) in the space Em. Any calcu-
lation in Em can also be carried out in Rn if the derivative of the vector c
is understood as the covariant derivative. For example, formulae (5.1) and
(5.2) should be written in the form

(B.12.16)

This means that if "we do not leave Rn", then we have to speak only of
that part of the increment in the vector c which can be determined in Rn.

B.13 Riemannian subspace Rn in the Euclidean


space En
The question of searching the Euclidean space Em which contains the Rie-
mannian space Rn determined by the square of the linear element (or the
covariant components of the metric tensor aa.(3) reduces to the search of
vector P (ql, ... ,qn) in Em from equalities (12.6) with prescribed right
B.13 Riemannian subspace R,. in the Euclidean space En 839

hand sides. The number of equations is ~n (n + 1) and the number of un-


known variables is m which is the number of projections of vector p on the
coordinate axes in Em, see eq. (12.1). If
1
m = 2"n(n+ 1),

then the above system of equations (6) has a real solution and this can be
proved by the methods of the theory of partial differential equations. For
instance, R2 is embedded in E3 whilst in general R3 is embedded only in
E 6 • However it may occur that Rn is embedded in an Euclidean space of
dimension m = n + p which is smaller than ~n (n + 1), then p is called the
class of R".
The curvilinear coordinates ql , ... ,qn, ... ,qn+p are introduced such that
the radius-vector p (ql, ... ,qn) of the points in E n+p belonging to Rn cor-
responds to the fixed values of the coordinates qn+ 1 , . .. ,qn+p , i.e.

p (q1, ... ,qn) -_ r (1


q , ... ,qn ,qon+l , ... ,qon+p ) . (B.13.1)

The base vectors in E n +p are denoted as


8r 8r
ra = 8qa' r n+s - --
- 8qn+s (a = 1, ... ,n; s = 1, ... ,p). (B.13.2)

The system of base vectors in Rn is given by

(B.13.3)

where zero denotes that the value defined in E n +p is considered in R". For
example, the values

gaf3 = ra· rf3, [a, (3 .,'Y1= ~ (8 ga ,


2 8 q f3 + 8g f3,
8 qa
_ 8 ga(3 )
8q, (B.13.4)

in R" are respectively equal to

( )0
gaf3 = aaf3, [a, (3,. 'Y10 -- 2"1 (8a a ,
8 q f3 + 8af3,
8qa -
8a a(3 ) - [
8q,
. 1
- a, (3, 'Y a .
(B.13.5)
Having the covariant components of the metric tensor in E n +p , i.e. the
matrix with the elements

we find the contravariant components which are the elements of the inverse
for this matrix
840 Appendix B. Basics of tensor calculus

With the help of these elements we find the vectors of the dual basis

(B.13.6)

where the summation over the dummy Greek and Latin indices is from 1 to
a and 1 to p respectively. It is seen from the latter equation that identifying
g~j3 with aaj3 orro with aaj3 Pj3 = pa is a grave error. Fully analogous, the
value

which is Christoffel's symbol of second kind defined in the metric of E n +p


but at points of R n , is not equal to these symbols calculated in the metric
of Rn.
It is known that the vectors r~+t are orthogonal to r a , hence, at points
of E n+p belonging to Rn we have
. rn+t = 0
Pa (B.13.7)
D'

i.e. r~+t are p vectors in space E n +p which are orthogonal to all vectors Pa ,
and thus orthogonal to Rn. It is obvious that any vector which is linearly
expressed in terms of r~+t is orthogonal to any P a . We can prove the
converse statement, namely, if m· P a = 0 for a = 1, ... ,n then vector m
is a linear form of vectors r~+t.
Taking into account the above, let us assume the vectors

(B.13.8)

are the base vectors of the space En+p" Let us express the second derivatives
r aj3 of the vector r in terms of these base vectors

(B.13.9)

The coefficients :f2, are not Christoffel's symbols of second kind. Indeed,
the scalar product of eq. (9) and rfj yields

(B.13.1O)

whereas the formulae in eq. (4.16) which define Christoffel's symbols of


second kind should take the form

It follows from eq. (10) that at the points in E n+p which belong to Rn we
have
B.13 Riemannian subspace Rn in the Euclidean space En 841

and this yields an important relationship

(B.13.11)

Let us proceed to determination of the coefficients qa,8ln+t. Calculating


the scalar product of eq. (9) and r n +s we arrive at the system of equations

qa,8ln+t g n+t,n+s -_ r a,8 . r n+s . (B.13.12)

Differentiating the relationships

with respect to q,8 and qa we obtain

Using eq. (4.19) we find

r a,8 . r n+s = ra . [{n(3+1 s} "( + {n


r (3 + s }
n+t r n+t] = {n(3+a s}
or

r a,8 . r n + s = gn+s,"( [a, (3; I] + gn+s,n+t [a, (3; n + t]. (B.13.13)

The system of equations (12) takes the form

Xn+t g n+t,n+s -- A "( gn+s,"( , (B.13.14)

where, for the sake of brevity,

Xn+t = qa,8ln+t - [a, (3; n + t] , A"( = [a, (3; I]· (B.13.15)

The quadratic form


n+p n+p
LL gkm pkPm = ga,8PaP,8 + 2ga,n+tpaPn+t + gn+t,n+sPn+tPn+s
k=l m=l

is the adjoint form of the quadratic form (dS)2, i.e. it is positive definite.
By virtue of Sylvester's theorem, the determinant
842 Appendix B. Basics of tensor calculus

of the system of equations (14) is positive. This means that the system has
a unique solution. Let us search it in the form

Xn+t = gn+t,,,Y"·

Then, noting that

9 n+t,,, g n+t,n+s + 9'1''' g'1',n+s = gn+s = 0


",

we obtain, instead of eq. (14), the following relationships

which can be easily satisfied if we define y" as the solution of the system
of equations

At the points of E n +p belonging to Rn this system has the form

and we immediately obtain that

yg = -a'1''' [a,.8; 1'la , x~+t = -a'1''' (gnH,,,)o [a,.8; 1'la . (B.13.16)

As mentioned above, the solution obtained is unique, hence

(B.13.17)

The values

(B.13.18)

are calculated in the following fashion. First, in the determinant

lal = (B.13.19)

the row l' is replaced by the following row

(B.13.20)

The ratio of the determinant obtained to determinant (19) is readily M;{+t.


In order to prove this, if is sufficient to represent the determinant obtained
as a sum of the products of the row (20) and the corresponding cofactors
B.13 Riemannian subspace Rn in the Euclidean space En 843

and to recall that lal a'Yv are the cofactors of the elements ofthe "f - th row
in the determinant (19).
Returning now to relationship (9) and accounting for formulae (11), (17)
and (18) we arrive at the following expressions for the second derivatives
r a{3 of the mdius-vector r at the points of Rn

Pa{3 = {a~} P'Y + Ua, /3; n+ tlo - MJ+t la, /3; "fla) r~+t
V' {3Pa + (la, /3; n + tl o - MJ+t la, /3; "fla) r~+t. (B.I3.2I)

Here

la, /3; n + tl =! (Oga,nH + og{3,n+t _ Oga(3) . (B.I3.22)


o 2 oq{3 oqa oqnH 0

Calculation of the coefficients of the expression of V' {3P a in terms of the


base vectors Pa and r~+t requires first knowledge of the values due to the
inner geometry of the surface. However, it is not sufficient as we also need
the values gn+t,a (t = 1, ... ,p; a = 1, ... , n) at points of Rn. The values
ga{3 (a,/3 = 1, ... ,n) are required in the vicinity of Rn belonging to E n +p
rather than in Rn, since la, /3; n + tlo contains the derivatives of ga{3 with
respect to the coordinate qn+t at points Rn, see eq. (22).
It is also important to notice that in the expression of a vector b in terms
of the base vectors (8) of the space E n +p

(B.I3.23)

the values lP and bn +t are neither contravariant nor covariant components


of b in the metric of En+p" Indeed, with the help of eqs. (2.2) and (2.3) we
can write

(B.I3.24)

At the points of E n +p belonging to Rn, we have

(B.I3.25)

so that the values ba at points of Rn are contravariant components of b in


the metric of Rn. Due to eq. (2.3), we can write
n+p
bn +q = L gn+q,sbs = gn+q,aba + gn+q,n+tbn+t ,
s=1

and the second set of relationships (24) takes the form

g n+q,aba -- gn+t,n+ q (b- n+t - bn+t·


) (B.I3.26)
844 Appendix B. Basics of tensor calculus

Again, we arrive at the system of equations (14) whose solution at points


of R n , by virtue of eqs. (16) and (17), is given by

(bnH ) 0 = (bnH)o - M~+t (boJa . (B.13.27)

Thus, the representation of the vector b, which does not belong to R n , at


points of Rn can be cast in the form
(B.13.28)
where (ba)a and (ba)a are the contravariant and covariant components of
the vector b in the metric of R n , respectively, and (bn+t)o is the value of
the covariant component of this vector at points of Rn.
Let us construct the expression for the derivative of a vector c given in
Rn. Denoting this derivative by b we obtain

b ac a a ac a
aqf3 = aqf3 c Pa = Pa aqf3 + Paf3c a (B
.13. 9
2)

Pa (~~: + {,8al' } ac'Y) + ([a,,8; n + tlo - M;[+t [a,,8; I'la) car~+t.


Turning to eq. (28) we see that the values
(ba)a = V' f3c a
are the contravariant components of b in the metric of Rn. Its covariant
components in the metric of E n+p at the point of Rn are given by
(bn+t)o = M~+t (ba)a + ca ([a,,8; n + tlo - M;[+t [a,,8; I'la) , (B.13.30)
where (ba)a are the covariant components in the metric of R n , i.e.
(B.13.31)
Nothing changes in the resulting formulae (12) and (30) if we consider
or n +s
embedding Rn into R n+p rather than into En+p' The vectors r af3' oqf3 '

:q~ should be replaced by the vectors V' f3r a, V' f3rn+s, V' f3c in the metric of
R n +p , as pointed out at the end of Sec. B.12.
In the particular case of R2 embedded into E 3 , i.e. for a surface in three-
dimensional space, we have

Ml = ~ I g~l g~2 I M2 - ~I an a12 I r30 = G


g033 m ,
3 lal a21 a22' 3 - lal g~l g~2' VYo-
where m is the unit vector of the normal to the surface. By means of the
determinant at the surface point
an a12 g~3
a21 a22 g~3
g~l g~2 g~3
B.14 The Riemann-Christoffel tensor 845

we find that

_ M1 = g31 Iglo = g3 1
3 0 lal g3 3 '

Formula (21) takes the form

Pof3 = {a~} /-Y + R (g3 [a, 3 (3; 3l o + gJ-Y [a, (3; 'Yla) . (B.13.32)

Comparing it with eq. (7.19) we again obtain expressions (9.16) for the
coefficients bga f3 of the second quadratic form of the surface.

B.14 The Riemann-Christoffel tensor


Let us consider two consequent covariant differentiations 'V f3 'V -y of the co-
variant values co. Recalling that 'V -yCO is a component of the tensor of
second rank which is covariant with respect to 'Y and contravariant with
respect to a and applying the rule of covariant differentiation, see last
formula in eq. (5.6), we obtain

It is easy to understand that the underlined terms appear in the expres-


sion for 'V f3 'V -yco. Thus, we obtain

(B.14.1)

where

The left hand side of relationship (1) contains the components of the
tensor of third rank which are covariant with respect to 'Y and (3 and con-
travariant with respect to a. Thus we can conclude that the values in eq.
(2) are the components of the tensor of fourth rank which are covariant
846 Appendix B. Basics of tensor calculus

with respect to {3, ,,/, p, and contravariant with respect to a. This tensor is
referred to as the Riemann-Christoffel tensor. Its covariant components are
expressed in a somewhat simpler form

(B.14.3)

It follows from eq. (1) and Ricci's theorem of Sec. B.5 that

(B.14.4)

The expression of the covariant components of the Riemann-Christoffel


tensor is transformed with the help of Christoffel's symbols of first kind

Then we have

g"'vo~i3gQO"b,p,;a]-g"'vO~1'gQO"[{3,p,;a] = 0~i3 b,p,;v]- O~1' [{3,p,;v]-


"'0' [ ] og",v
9 ,,/, p,; a oqi3 + 9"'0' [{3 , p,; a ] og",v
oq1'·

Expanding the expressions for the derivatives of the brackets and substi-
tuting the following equalities

;q"'; = [a, {3; v] + [{3, v; a] , o;",v = [a,,,/;v]


uq1'
+ b,v;a],
we obtain
1 ( 02g1'V 0291'11- 029,1311- 02gi3v )
'2 oqi3 oql1- - oqi3 oqV + oq1' oqV - oq1' oql1- -
g"'O' (b, p,; a] [v, {3; a]- [{3, p,; a] [v, "/; a]) . (B.14.5)

From the total number of n 4 of the components of the Riemann-Christoffel


tensor there are An2 (n - 1)2 independent components. For instance, for
n = 2 there exists a single independent component R1212 whilst for n = 3
we have six independent components.
As mentioned in Sec. B.ll the distinctive feature of Euclidean space
is the feasibility of introducing a (Cartesian or non-orthogonal) coordi-
nate system with axes of fixed directions for which the expression for the
quadratic form has constant coefficients. In particular, it is simply a sum
of squares for a Cartesian coordinate system. But then all Christoffel's
symbols are zero, as well as all the components of the Riemann-Christoffel
tensor. This tensor property can not be bound to the choice of the coor-
dinate system. Naturally, the converse statement can be proved, namely, if
B.14 The Riemann-Christoffel tensor 847

the Riemann-Christoffel tensor is identically equal to zero, then the man-


ifold is Euclidean. Alternatively, there can be assumed such coordinates
ql, ... ,if' that the quadratic form (ds)2 , expressed in terms of these, will
have constant coefficients. The non-trivial Riemann-Christoffel tensor tes-
tifies that the manifold under consideration is not Euclidean. If (ds)2 is
a positive definite form of the differentials dqk then it is a Riemannian
manifold Rn.
Therefore, for any choice of the curvilinear coordinates in En

(B.14.6)

i.e. the sequence of covariant differentiation can be changed.


References

[1] Appell, P. Traite de Mecanique rationelle. Paris, Gauthier-Villars,


1926.

[2] Appell, P. Sur une forme generale des equations de la dynamique.


Memorial des Sc. Math. Paris, No.1, 1925.

[3] Babakov, I.M. Theory of vibration (in Russian). Gostekhizdat, 1958.

[4] Beletsky, V.V. Motion of an Earth satellite about its centre of mass
(in Russian). In: Earth's satellites, The USSR Academy of Science,
VoLl, 25-43.

[5] Beyer, R Dynamik der Mehrkurbelgetriebe. Zeitschrift fur Ange-


wandte Mathematik und Mechanik, 8, No.2, 1928.

[6] Biezeno, C.B., Grammel, R Technische Dynamik. Springer, Berlin,


1953.

[7] Bisplinghoff, RL., Ashley, H., Halfman, RL. Aeroelasticity. Addison-


Wesley, Cambridge, 1955.

[8] Blitzer, L., Weisfeld, M., Wheelon, A. Perturbation of the orbit of a


satellite, due to the Earth's oblateness. Journal of Applied Physics,
vol. 27, No. 10, 1956, 1141-1149.

[9] Blitzer, L. Effect of the Earth's oblateness on the period of a satellite.


Jet propulsion, April 1957, 405-406.
850 References

[10] Bobylev, D.K. On a sphere with a gyroscope inside it (in Russian).


Collection of Mathematical Papers, vol. 16, 1892.

[11] Bobylev, D.K. On the principle of Hamilton or Ostrogradsky and the


principle of least action (in Russian). Supplement to Volume LXI of
the Russian Academy of Science, 1899.

[12] Bogolyubov, N.N., Mitropolsky, Yu. A. Asymptotic methods in the


theory of nonlinear vibrations (in Russian). Fizmatgiz, 1958.

[13] Boltzmann, L. Uber die Form der Lagrange'schen Gleichungen


fur nichtholonome generalisierte Koordinaten. Sitzungsberichte der
Wiener Akademie der Wissenschaften, B. 111, Dez. 1902.

[14] Bulgakov, B.V. Applied theory of gyroscopes (in Russian). Gostekhiz-


dat, 1955.

[15] Bulgakov, B.V. Vibrations (in Russian). Gostekhizdat, 1954.

[16] Caratheodory, C. Variationsrechnung und partielle Differentialgle-


ichungen erster Ordnung, B. 1, Leipzig, 1956.

[17] Chaplygin, S.A. On the motion of a heavy rigid body of revolution


on a horizontal plane (in Russian), 1897. In: Chapygin, S.A. Investi-
gations of dynamics of non-holonomic systems. Gostekhizdat, 1949.

[18] Chaplygin, S.A. On a sphere rolling on a horizontal plane (in Rus-


sian), 1903. In: Chapygin, S.A. Investigations of dynamics of non-
holonomic systems. Gostekhizdat, 1949.

[19] Charlier, C.L. Die Mechanik des Himmels, Leipzig, 1902.

[20] Chetaev, N.G. Stability of motion (in Russian). Gostekhizdat, 1955.

[21] Chetaev, N.G. On a gyroscope in Cardan's suspension (in Russian).


Applied Mathematics and Mechanics, vol. 22, 1958, 379-38l.

[22] Dzhanelidze, G. Yu., Lurie, A.I. On the application of the integral


and variational principles of mechanics to vibrational problems (in
Russian). Applied Mathematics and Mechanics, vol. 24, 1960, 80-87.

[23] Darboux, G. Le<;ons sur la tMorie generale des surfaces. T.1, paris,
1887.

[24] Duboshin,G.N. Introduction to celestial mechanics (in Russian).


ONTI,1938.

[25] Fowler, R.H., Gallop, E.G., Lock, C.H. and Richmond, H.W. The
aerodynamics of a spinning shell. Philosophical Transactions of the
Royal Society of London, 221-222, 1921-1922.
References 851

[26] Frazer, R.A., Duncan, W.J., Collar, A.R. Elementary matrices and
some applications to dynamics and differential equations. Cambridge,
The University Press, 1938.

[27] Gantmakher, F.R. Theory of matrices (in Russian). Gostekhizdat,


1953.

[28] Gantmakher, F.R., Levin, A.M. On the equations of motion of a


rocket (in Russian). Applied Mathematics and Mechanics, vol. 11,
1947, 301-312.

[29] Gantmakher, F.R., Levin, A.M. Theory of flight of uncontrolled rock-


ets (in Russian). Fizmatgiz, 1959.

[30] Golubev, V.V. Lectures on integration of the equations of motion of


a heavy rigid body about a fixed point (in Russian). GTTI, 1953.

[31] Grammel, R. Der Kreisel. Berlin, Springer, 1950.

[32] Grinberg, G.A. Selected problems of focusing electron beams (in Rus-
sian). Publishers of the USSR Academy of Science, 1948.

[33] Goursat, E.J.-B. Cours d'Analyse. Paris, 1926.

[34] Gyunter, N.M. A course of variational calculus (in Russian).


Gostekhizdat, 1941.

[35] Hamel, G. Die Lagrange-Eulerschen Gleichungen der Mechanik.


Zeitschrift fur Mathematik und Physik, B. 50, 1904.

[36] Hamel, G. Theoretische Mechanik, Berlin, Springer, 1949.

[37] Hertz, H. Die Prinzipien der Mechanik in neuen Zusammenhange


dargestellt. 1894.

[38] Heun, K. Lehrbuch der Mechanik, Teil 1, Kinematik, Sammlung


Schubert, Leipzig, 1906, S. 133-153.

[39] Holder, O. Gottinger Nachrichten, 1896, p.112.

[40] Idelson, N.!. Theory of potential and its application to geophysical


problems (in Russian). GTTI, 1932.

[41] Isaeva, L.S. On the sufficient conditions of stability of a heavy top


on a rough plane (in Russian). Applied Mathematics and Mechanics,
vol. 23, 1959, 403-406.

[42] Ishlinsky, A.Yu. Mechanics of special gyroscopic systems (in Russian).


Publishers of the Ukrainian Academy of Science, Kiev, 1952.
852 References

[43J Ishlinsky, A.Yu. On the theory of complex systems of gyroscopic sta-


bilisation (in Russian). Applied Mathematics and Mechanics, vol. 22,
1958,359.

[44J Jacobi, C.G.J. Vorlesungen tiber Dynamik. Berlin, 1884.

[45J Jordan, C. Cours d'analyse de l'Ecole Poly technique, t.3, Paris,


Gauthier-Villars, 1887.

[46J Kantorovich, L.V., Krylov, Approximate methods of higher analysis


(in Russian). Gostekhizdat, 1941.

[47J Kilchevsky, N.A. Elements of tensor calculus and its application in


mechanics (in Russian). Gostekhizdat, 1954.

[48J Kirchhoff, G. Vorlesungen tiber mathematische Physik, Mechanik.


Leipzig, 1876.

[49J Klein, F. Vorlesungen tiber die Entwicklung der Mathematik im 19.


Jahrhundert. Springer, Berlin 1927.

[50J Klein F. und Sommerfeld, A. Uber die Theorie des Kreisels. Heft 1,
Kap. 1, S. 7-68, Leipzig, 1897.

[51J Kochin, N.E. Vector calculus and the basics of tensor calculus (in
Russian). GTTI, 1938.

[52J Krylov, A.N. Sur la variation des elements des orbites elliptique de
planetes, 1915. In: Krylov, A.N Collection of papers. Publishers of
the USSR Academy of Science, 1955, vol. 6. 249-266.

[53J Krylov, A.N. Cardan's suspension in ships (in Russian). In: Krylov,
A.N Collection of papers. Publishers of the USSR Academy of Sci-
ence, 1955, voLl2, 193-258.

[54J Krylov, N.M., Bogolybov, N.N. Introduction to nonlinear mechanics


(in Russian). Kiev, 1937.

[55J Lagrange, J.1. Mechanique Analytique. Paris, 1788.

[56J Loizansky, L.G. and Lurie A.I. A course of Theoretical Mechanics (in
Russian). Gostekhizdat, 1955.

[57J Lurie A.I. On the theory of finite rotation of a rigid body (in Russian).
Applied Mathematics and Mechanics, vol. 21, 1957, 371-373.

[58J Lurie A.I. Notes on analytical mechanics (in Russian). Applied Math-
ematics and Mechanics, vol. 21, 1957, No.6.
References 853

[59] Lurie, A.I. Equations of perturbed motion in the Keplerian problem


(in Russian). Applied Mathematics and Mechanics, vol. 23, 1959,
412-413.

[60] Lurie, A.I. Some problems of dynamics of rigid body systems (in
Russian). Transactions of the Leningrad Poly technical Institute 1960,
vol. 210, 7-22.

[61] Magnus, K. On the stability of motion of a symmetric gyroscope


in Cardan's suspension (in Russian). Applied Mathematics and Me-
chanics, vol. 22, 1958, 173-178.

[62] Malkin, I.G. Theory of motion stability (in Russian). Gostekhizdat,


1952.

[63] Maltsev, A.I. Basics of linear algebra (in Russian). Gostekhizdat,


1955.

[64] Maupertius, P.L.N. de. Mem. de l'Acad., 1744, p.417.

[65] McMillan, W.D. Dynamics of rigid bodies. New York, Dower, 1960.

[66] Merkin, D.R. Gyroscopic systems (in Russian). Gostekhizdat, 1956.

[67] Mikhlin, S.G. Direct methods in mathematical physics (in Russian).


Gostekhizdat, 1950.

[68] Nielsen, K.L. and Synge, J.L. On the motion of a spinning shell.
Quarterly of Applied Mathematics, IV, No.3, 1946.

[69] Nikolai E.L. On the theory of deviation of gyroscopic devices (in


Russian). In: Nikolai E.L., Works on mechanics, Gostekhizdat, 1955.

[70] Nikolai E.L. On the motion of a balanced gyroscope in Cardan's


suspension (in Russian), 1939. In: Nikolai E.L., Works on mechanics,
Gostekhizdat, 1955, 455-496.

[71] Nikolenko, G.I. Vibrations of prestressed elastic systems (in Russian).


Collection of papers in Engineering, vol.l1, 1952, 79-94.

[72] Novoselov, V.S. An example of nonlinear non-holonomic constraint


which is not reducable to Chetaev's type (in Russian). Transactions
of the Leningrad University, Vol. 19, 1957, 106-112.

[73] Okhotsimsky, D.E., Eneev, T.M., Taratynova, G.P. Determination


of the time expectancy of an Earth satellite and an investigation of
secular perturbations of its orbit (in Russian). Successes of Physical
Sciences, vol. 63, 1957, 33-50.
854 References

[74] Polak, L.S. ed. Variational principles of mechanics (in Russian). Fiz-
matgiz, 1959.
[75] Rayleigh, Lord Strutt The theory of sound. 1896.
[76] Roberson, R.E. Torques on a satellite vehicle from internal moving
parts. Journal of Applied Mechanics, vol. 25. No.2, 196-200, 1958.
[77] Roze, N.V. Lectures on analytical mechanics (in Russian). Publishers
of the Leningrad University, 1938.
[78] Rumyantsev, V.V. On the stability of motion of a symmetric gyro-
scope in Cardan's suspension (in Russian). Applied Mathematics and
Mechanics, vol. 22, 1958, 374-38l.
[79] Rumyantsev, V.V. On the stability of motion of a symmetric gyro-
scope in Cardan's suspension, part II (in Russian). Applied Mathe-
matics and Mechanics, vol. 22, 1958, 499-503.
[80] Saxon, D., Cahn, A.S. Modes of vibration of a suspended chain. Jour-
nal of Mechanics and Applied Mathematics, vol. 6, part 3, 1953.
[81] Shubenko, L.A. On the influence of centrifugal forces on the frequency
of free vibration of blades of steam turbines (in Russian). Transac-
tions of the Leningrad Industrial Institute, 1937, No.6, 106-112.
[82] Stepanov. A course on differential equations (in Russian). GITTL,
1945.
[83] Stiickler, R. Uber die Berechnung der an rollenden Fahrzeugen wirk-
enden Haftreibungen. Ingenieur-Archiv, 1955, 23, No.4, 279-287.
[84] Sushkevich, A.K. Basics of linear algebra (in Russian). ONTI, 1937.
[85] Suslov, G.K. Theoretical mechanics (in Russian). Gostekhizdat, 1944.
[86] Synge, J.L. On the geometry of dynamics. Philosophical Transactions
of the Royal Society of London, Ser. A.,Vol. 226, 31-106.
[87] Taratynova, G.P. On the motion of a satellite in a non-central field
of attraction of the Earth and the influence of atmospheric resistance
(in Russian). Successes of Physical Sciences, vol. 63, 1957, 50-58.
[88] Thomson, W., Tait, P.G. Treatise on Natural Philosophy. VoLl, 1879.
[89] Tisserand, F. Traite de mechanique celeste, 2, Paris, 189l.
[90] Vallee Poussin, C.J.G.N. de la, Cours d'Analyse. Paris, 1909.
[91] Voronetz, P.V. Equations of motion of a rigid body rolling On a fixed
plane without slipping (in Russian). Kiev, 1903.
References 855

[92] Vyshnegradsky, LA. On direct-acting controllers (in Russian). In:


Automatic control, collection of papers. Publishers of the USSR
Academy of Science, 1949, 43-87.

[93] Webster, A.G. The dynamics of particles and of rigid, elastic, and
fluid bodies. Leipzig, B.G. Teubner, 1904.

[94] Whittaker, E.T. and Watson, G.N. A course of Modern Analysis.


Cambridge University Press, London, 1935.

[95] Whittaker, E.T. A Treatise on the Analytical Dynamics of Particles


and Rigid Bodies. Cambridge University Press, Cambridge, 1937.

[96] Yatsunsky, LM. On the influence of geophysical factors on motion of


a satellite (in Russian). Successes of Physical Sciences, vol. 63, 1957,
59-71.

[97] Zhukovsky, N.E. On strength of the motion (in Russian), 1882. In:
Zhukovsky, N.E., Collection of papers, Gostekhizdat, 1937, 110-208.
[98] Zhukovsky, N.E. On the motion of a rigid body with cavities filled by
homogenous fluid (in Russian), 1885. In: Zhukovsky, N.E., Collection
of papers, Gostekhizdat, 1936, 21-186.
[99] Zhukovsky, N.E. On the gyroscopic sphere of Bobylev (in Russian),
1894. In: Zhukovsky, N.E., Collection of papers, Gostekhizdat, 1948,
352.

[100] Zhukovsky, N.E. Condition for equilibrium of a rigid body on a


plane with consideration of the friction force (in Russian), 1896. In:
Zhukovsky, N.E., Collection of papers, Gostekhizdat, 1937,433-449.

[101] Zhukovsky, N.E. Collection of papers. Lectures (in Russian).


Gostekhizdat, 1939, 45-48.
Index

acceleration precession, 52, 54


absolute, 89 roll, 55
of a particle moving over slide, 55
the rotating earth, 93 spin, 52, 54
addition, 90 trim, 58
angular, 81 velocity roll, 55
centripetal, 82 yaw, 55, 58
Cariolis, 90 angles
generalised, 26 Euler's, 51, 54
of a point in a rigid body, 81, angular momentum
86 relative, 173
relative, 89, 90 angular velocity, 130, 138
rotational, 82 Darboux's equation, 142
translational, 89, 90 determination of a rigid body
of a particle moving over position, 139, 142
the rotating earth, 92 anomaly
action eccentric, 578
Lagrange's, 737, 776 mean, 579
angle true, 579
ascend,55 aphelion, 577
attack, 55 apocentron, 577
heading, 55 apogee, 577
heel, 58 axes
nutation, 52, 54 airplane, 54
pitch, 55 ship, 58
858 Index

velocity, 55 coordinate
cyclic, 371
base vector, 805, 819 generalised, 23
binormal, 106 cyclic, 367
bracket excessive, 28
Lagrange, 539, 785 variation, 34, 36
fundamental, 539 ignorable, 371
Poisson, 537, 786 positional, 367, 371
fundamental, 538 quasi-cyclic, 388
transformation, 50
canonical transformation coordinates
invariance, 550 curvilinear orthogonal, 828
Cardan's suspension, 62, 67, 125, curvature, 106
129 line, 823
double, 64 normal section of surface, 823
Cayley-Klein's parameters, 135, 136 surface line
centre of inertia, 158 geodesic, 824
chain hanging with a mass on the
end, 698, 723 degrees of freedom, 22
Christoffel's square brackets, 177 number, 22
Christoffel's symbol derivative
first kind, 177 covariant, 816
second kind, 312, 813 on surface, 827
coefficients derivative of vector
gyroscopic, 310 relative (local), 87
Lame, 828 determinant, 783
composition of motions, 102 differential of quasi-coordinates, 32
condition of kinematic feasibility differentiation
of adjacent motion, 692 covariant, 837
constraint, 19 in Riemannian space, 837
force of, 267 matrix, 802
holonomic, 19 of surface vector, 825
non-stationary, 21 discriminant of quadratic form, 791
rheonomic, 21 displacement
scleronomic, 21 virtual, 35
stationary, 21 double-crank mechanism, 28
ideal, 271 dual vector, 805, 820
non-holonomic, 20 dyad, 158
one-sided, 20
principle of constraint release, elastic body, 227
267 element
two-sided, 20 kinematic, 742
constraint force, 267 metric,742
generalised, 349, 353, 358 elements
constraint release, 267 elliptic orbital, 581
Index 859

energy of system of particles, 278


accelerations, 176 stable, 280
kinetic, 151 unstable, 280
amplitude, 718 relative, 461
in terms of generalised ve- Euclidean subspace, 838
locities, 151, 152, 171 Euler's angles, 51, 54
in terms of quasi-velocities,
155,171 finite rotation, 114, 130
of system subjected to sta- commutation, 122
tionary constraints, 153 composition, 120
potential, 208, 209 in terms of Euler's angles, 124
amplitude, 718 Rodrigues-Hamilton's param-
generalised, 211 eters, 115
of attraction force, 217 subtraction, 122
energy integral, 308 foci
equation conjugate kinetic, 677, 691
central force
fundamental, 275 aerodynamic resisting, 260
Lagrange's, 275 circulatory, 215
Darboux-Rikatti, 143 Coriolis's, 514
Kepler, 579 dissipative, 248
of dynamics, fundamental, 273 power, 251
relative motion elastic, 227
canonical, 459 generalised
equations for generalised coordinate,
canonical 203
Hamiltonian, 529, 590 for quasi-coordinate, 204,
of relative motion, 555 207
variational, 631, 633 in terms of the potential
integration, 634 energy, 209
equations of motion gravity, 216
Appell's, 418, 420, 423 gyroscopic, 310
Chaplygin's, 424 generalised, 310
Euler-Lagrange, 392 positional, 213, 214
Lagrange's potential, 208, 211
of the first kind, 270 indication, 210
perturbed,586,623,637,649 radial correction, 215
Routh, 371 resisting, 250
the Euler-Lagrange, 434 thrust, 514
equilibrium formula
in presence of Coulomb's fric- Frenet, 107,823
tion, 296 friction
of a rigid body suspended on Coulomb, 249, 254, 296
elastic rods, 291 viscous, 249
of heavy rod, 283 friction gear, 429
860 Index

frictional pole, 299 perturbed, 625, 642


function kinetic energy
dissipation, 249, 251 in terms of generalised veloc-
generating, 523, 546 ities, 171
Hamilton's in terms of quasi-velocities,
characteristic, 767, 769 171
principal, 730, 731 of relative motion, 175
Hamiltonian, 528 of translational motion, 175
homogeneous, 154 kinetic potential
Lagrange, 276 Routhian, 372
functional Kronecker symbol, 48
stationary value, 672
Lagrange's equations
generalised momentum, 155 of the first art, 270
geodesic line, 823, 825 of the second kind, 304
gravitational field, 216, 288 Lagrange's multipliers, 269
gyroscope Legendre's condition, 686
in Cardan's suspension, 183, Levi-Civita symbol, 48
334
mounted on a moving plat- Magnus effect, 262
form, 468 mathematical pendulum, 250
gyroscopic platform, 188, 515 double
with a square-law resisting
Hamilton's action, 670, 679 force, 252, 329
minimum, 676 matrix
variation, 671 addition, 787
heavy top adjoint, 794
influence of the rotating Earth, block, 784
484 multiplication, 792
helical spring, 245 conform, 788
Hooke's law, 228 deficiency, 784
determinant, 783
ideal constraint
diagonal elements, 786
constraint force, 356
differentiation, 802
indication of potential force, 210
Hess's, 783
inertia ellipsoid, 168
influence, 231
central, 168
inverse, 795
inertia tensor, 161, 163
Jacobi's, 782
transformation, 162
multiplication, 788
Jacobi's condition, 686 by scalar, 787
Jacobi's identity, 538 null, 782
Jacobian, 784 of inertia, 163
order, 782
Keplerian motion, 377 orthogonal, 800
Earth satellite Poisson's, 540, 786
Index 861

rectangular, 781 elliptic, 582


rotation, 50 in a resistive medium, 598
square, 782 in homogeneous gravitational
skew-symmetric, 786 field, 770, 778
symmetric, 786 on a conical surface, 757
subtraction, 787 on developable surface, 325
transpose, 786 on surface, 318
unit, 783 on surface of revolution, 324
mean motion, 579 perturbed, 637
mechanism, 360 planar, 428
crankshaft, 363 relative, 88
method rigid body
Galerkin's, 723 carrying rotating flywheeb,
ignoration of cyclic coordinates, 486
371 free, 438, 440, 442
Ritz's, 719 having a cavity with fluid,
variation of parameters, 585 492
metric relative, 475
element of action, 742 rocket, 512
moment of inertia, 161 spinning shell, 260, 443
about an axis, 161 stability, 747
principal central, 168 unperturbed stationary, 660
momentum, 172 kinematically stable, 660
motion
Earth satellite, 625, 642 nodal axis, 52
elastic solid, 498
gyroscopic platform, 515, 519 parallel translation of vector, 826
heavy top, 533 parameters of Rodrigues-Hamilton,
carrying a flywheel, 536 115
influence of misbalance, 605 path
in field true, 670, 742
of a central force, 377, 532, varied, 670, 682, 742
552, 567, 568, 576, 756, pendulum
772 mathematical, 727
of attraction of two cen- physical, 356
tres, 760 pericentron, 577
in the gravity field of the ro- perigee, 577
tating Earth, 590 perihelion, 577
Keplerian, 377, 532, 552, 567, perturbation of trajectories, 747
568, 576, 756, 772 plane
perturbed, 621 tangent contact, 105
material system Poisson bracket, 537
perturbed, 649 potential
natural trihedron, 105 kinetic, 276
particle power, 247
862 Index

dissipative forces, 251 rod


in terms of generalised veloc- elastic rotating, 707, 726
ities, 248 inextensible, 506
virtual, 248 heavy, 283
principal angular momentum, 172 in an elliptic cup, 286
principal axes of inertia, 165 plain axis, 242
central, 168 under bending, torsion and
principal normal to a curve, 106 compression, 238
principle rod system
Hamilton-Ostrogradsky, 672, statically determinate, 231
698 statically indeterminate, 234
of constraint release, 267 Rodrigues formula, 112
of virtual work, 280 rolling, 402
stationary action, of without slipping, 93
Jacobi, 740 ring, 79
Lagrange, 737 sphere, 79
product rotation
dyadic, 158, 789 about a fixed point, 61
scalar, 834 earth satellite about its cen-
product of inertia, 161 tre of inertia, 611, 620
flexible shaft, 462
quadratic form, 797 rigid body, 147
adjoint, 796 about a fixed point, 170,
of surface 180
first, 821 near vertical, 761, 765
second, 821 with a fixed point, 483
positive Routhian function, 371, 373
definite, 797
semi-definite, 797 safety parabola, 756, 779
quasi-coordinates, 32 shape of the earth, 221
quasi-velocities, 30 shell carrying flywheels
kinetic energy, 184
regular precession, 147, 665 solid,498
resultant force, 205 space
rigid body, 19 Euclidean, 833
carrying an unbalanced fly- Riemannian, 836
wheel, 185 sphere
in a central force field, 288 rolling on a rough surface, 193,
suspended on elastic rods, 291 395, 425
with a fixed point, 170, 213 spinning shell, 443
rigidity stability
bending, 240 of regular precession of gyro-
torsional, 240 scope, 665
ring stability of motion, 747
rolling on a rough surface, 425 submatrix, 783
Index 863

Sylvester's criterion, 153 generalised, 525


symbol Euler-Chasles, 117
Christofell's Jacobi's, 562
first kind, 813 Lagrange-Dirichlet, 280
second kind, 813 Liouville's, 549
Levi-Civita, 807 Meusnier, 823
symbols Poisson, 541
three-index, 39, 40 reciprocal Maxwell's, 231
system Ricci, 817
four-rod,343 Stackel's, 571
free, 19 Steiner's, 163
holonomic, 20 Sylvester's, 798
linear equations theory
solution, 796 Sommerfeld's, 216
Liouville, 574 three-index symbols, 76
particle, 19 trajectory of a particle in the po-
prestressed, 292 tential field, 316
three-rod, 281 transformation
two heavy tops, 383 canonical, 543, 559
two-rod, 339, 364 Legendre's, 523, 524
system of axes trihedron
geocentric, 91 natural, 106
rotation, 127 motion, 109
inertial, 47 twist of curve, 107
rotation matrix, 50 two-axle trolley, 42, 196, 400, 425
moving, 47
rotation matrix, 50 variables
canonical, 529
tangent vector, 105 variation
tensor, 802, 834 asynchronous, 734
dyadic representation, 810, 834 coordinate
fundamental, 810 synchronous, 36
in non-orthogonal coordinates, generalised coordinate, 34
810 quasi-coordinate, 39
metric, 812, 820 synchronous, 734
of second rank vector
symmetric, 160 angular acceleration, 81
Ricci, 654 angular velocity, 71, 74
Riemann-Christoffel, 846 components
skew-symmetric, 811 contravariant, 808
symmetric, 810 covariant, 808
theorem Darboux, 107
addition of accelerations, 90 geodesic curvature, 823
addition of velocities, 89 in Riemannian space, 837
Castigliano, 231 infinitesimal rotation, 73
864 Index

of finite rotation, 114, 115,


120, 122, 130
vector basis, 805
dual, 805
vector of momentum, 172
relative, 173
translational, 173
vector system
half-fixed, 52
half-movable, 52
velocity
absolute, 89
of point moving over the
rotating earth, 91
addition, 89
angular, 71
generalised, 25
of a point in a rigid body, 71,
84
matrix form, 84
relative, 89
translational, 89
of point moving over the
rotating earth, 91
virtual, 26
vibration
about the equilibrium, 599
chain line, 710
hanging chain with a mass,
698, 723
mathematical pendulum, 250,
252, 329, 727
particles attached to a mov-
ing rigid shell, 490
physical pendulum, 356
rotating rod, 506, 707, 726

work
complementary, 526
elementary, 203, 204, 208
of potential forces, 209

You might also like