Вы находитесь на странице: 1из 858

Foundations of Engineering Mechanics A.

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

Analytical Mechanics

Translated by A. Belyaev

With 92 Figures
Series Editors:
Vladimir I. Babitsky J. Wittenberg
Department of Mechanical Engineering Institut fur Technische Mechanik
Loughborough University Universitat Karlsruhe (TH)
LE11 3TU Loughborough, Leicestershire Kaiserstrafie 12 76128 Karlsruhe /
Great Britain Germany

Author:
A. I. 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 bibliographical references and index.
ISBN 978-3-642-53650-2

This work is subject to copyright. All 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 publishing group


http://www.springer.de
© Springer-Verlag Berlin Heidelberg 2002
Originally published by Springer-Verlag Berlin Heidelberg
New York 2002 Softcover reprint of the hardcover 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 regulations and
therefore free for general use.

Typesetting: Camera-ready copy from authors Cover-


Design: de’blik, Berlin
Printed on acid-free paper SPIN 10859752 62/3020/kk 5 4 3 2 1 0
Anatolii I. Lurie

This book was written by a great Russian 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 graduated from a
high school (gymnasium), and was admitted to the Faculty of Physics and Mechanics of
the Saint-Petersburg Polytechnic Institute, 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
the Department, 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 the Department of Mechanics and Control Processes in
1960). A.I. Lurie was a Corresponding Member of the USSR Academy of Sciences,
Division of Mechanics and Control Processes. He was a member of the Presidium of the
National Committee for Theoretical and Applied Mechanics and a member of the National
Committee for Automatic Control. A.I. Lurie was a member of the Editorial Boards of the
renowned Russian journals ’’Applied Mathematics and Mechanics”, and ’’Mechanics of
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.I. Theoretical Mechanics. In three volumes.
Leningrad, Moscow, GMTI, 1934.
3. Lurie, A.I. Statics of Thin-walled Elastic Shells. Moscow, Gostekhiz- dat, 1947,
252 pp.
4

4. Lurie, A.I. Some Nonlinear Problems of the Theory of


Automatic Control. Moscow, Gostekhizdat, 1951, 216 pp.
5. Lurie, A.I. Operational Calculus and its Application to the
Problems in Mechanics. Moscow, GITTL, 1951, 432 pp.
6. Lurie, A.I. Three-dimensional Problems of the Theory of
Elasticity. Moscow, GITTL, 1955. 492 pp.
7. Loitsianskii, L.G. and Lurie, A.I. 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
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): ”1
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 ”

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, variational 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 subjects which, in the author’s opinion, are most
closely related to engineering problems.
Application of the methods of analytical mechanics to non-
trivial problems at the very stage of constructing the equations
requires detailed knowledge of the issues that are normally
only briefly touched upon. With this perspective considerable
attention is paid to ways of introducing the generalised
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 problems of analytical
statics are discussed as well. It is thought useful to include
8

This is followed by the study of canonical equations, canonical


transformation and the problem 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 problem 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 indicates 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 Polytechnic Institute1. However
the author hopes that students and researchers in various
fields of engineering will find this book useful.

1
Translator’s note: now the State Technical University of St. Petersburg
Translator’s preface

The book ” Analytical Mechanics” by A.I. 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 Western 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.
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....Ill
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
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
14 Contents

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


6 The fundamental equation of dynamics. Analytical statics
267
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 central 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
Contents 15

7.11.2......Plane
motion of a heavy rigid body on a
string passing 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 Cycliccoordinates................................................. 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 Ring...............................................................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 Ring...............................................................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
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’stheorem 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
carrying 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
lO.lOExamples of canonical transformations.................552
10.10.1 First example...............................................552
10.10.2 Second example.........................................553
10.11 Canonical equations of the relative motion..........554
10.12Canonical transformation and the process of motion
. . . . 557
Contents 17

10.14.5Liouville’s system........................................574
10.15Keplerian 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
11.10Variational equations............................................631
11.11 On 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
11.17St at ionary unperturbed motions ........................660
11.18 Examples.............................................................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
18 Contents

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


12.11 Jacobi’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
12.14.4 Motion 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
12.16.2Keplerian motion.........................................772
12.170n the character of the extremum of Lagrange’s
action . . . 776
A Elements of the theory of matrices 781
A.l 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. l 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
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 OMi is denoted by r*, when the subscript i is 1,
2, ...A, and N is the number of particles in the system.
The simplest and most important class of constraints are
holonomic constraints. These ensure dependences between the
coordinates of the system’s points and are expressed
analytically in terms of following relations
fi (x1,y1,z1, ...,xN,yN,zN;t) = 0 (i = 1, ...,r), (1.1.1)
20 1. Basic definitions

which are called constraint equations1. The number of


constraint equations is denoted by r. It is clear that r < 37V,
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 Z, the second end of which is
fixed at the coordinate system origin, which is subject to the
Z2 - (x2 + y2 + z2) > 0,

indicating that the distance between particle M and the


coordinate system origin does not exceed Z. Constraints of
this sort are referred to as onesided 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 consideration here to non-holonomic
constraints which correspond to linear projections of the
particle velocities. The constraint equations are
N

where a^, bik and depend on the coordinates of the particles


and time. Equation (2) is equivalent to the following
N
Y, (aikdxk 4- bikdyk + cikdzk) + gidt = 0. (1.1.3)
fc=1

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
Chapter, indicates
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
daik dais daik dbis daik _ dcis daik _ dgi '
dxs dxk ’ dys ~ dxk ’ dzs dxk ’ dt dxk’
dbik _ dbis dhk dcis dhk _ dgi ^
dys dyk ’ dzs dyk ’ dt dyk’ *
dcik _ dcis d^k _ dgi

dzs dt dzk ’ ,
dzk
are met ■ all k and s from 1 a fixed , the i-th (1.1.4)
for to . N for i equation
integrabl as all ; in can be satisfied by casting
e, equalities (4)
dfi , _dfi_ = dfj_ dfi
ttik —TJ ’ °ik dyk dzk 9l dt ’
dxk

(1.1.5)

where fi is a function of coordinate and time. Under this


condition eq. (3) expresses that a perfect differential dfi is
equal to zero, i.e.
fi (xi,yi,zi, ...,xN,yN,zN;t) - Ct = 0. (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 a^, bik and in eq. (2)
do not depend explicitly on time, and in addition to this gi = 0.
When ^ / 0 a constraint is considered as rheonomic because t
is present in eq. (3) via dt even though factors a^, 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
fi (xi,yi, Z!, ...,XN,yN, zN) - git-Ci = 0, (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
x
v — fi 3*y-2> Vv ~ ^3^-15 z
v — fi?>v (r = 1 ,...,iV).
dfi dfr
" d£ 1
Df
D£ dfi dfr
d^3N d^3N

in the domain of the variables £1?..., £3iV for all time t must be
equal to r, i.e. the deficiency of matrix (4) is zero. For
instance, let the Jacobian
dfl dfr
dZi ■" dh
(1.2.5)
dfl Ofr
d^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+1, ...,£3N an(i time t. Thus we obtain
relations of the form
Zk = tk (tr+l,-,t3N',t) (k = l,...,r), (1.2.6)
in which coordinates £r+1, ...,£3N are independent of each other. If
a system 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 coordinates £1, ...,£r are determined in terms
of the independent coordinates by virtue of eq. (6). Given non-
holonomic constraints, the number of independent
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 coordinates £r+1, ...,£3iV. Instead of these
one can introduce any other independent quantities qi,...,qn =
2
The 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
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
£r+fc = tr+k(qU-,Qn;t) (k = 1, ...,Tl) . (1.2.7)
The condition for this is a non-zero
Jacobian
d£r+1 d£r+1
dqi dqn 7^ 0, (1.2.8)
Ji= .......................................
dqi dqn
expressing independence of the quantities and solvability of
equation
(7) for <?i,...,gn.
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
gi,...,gn referred to as the generalised coordinates and time t.
Returning
to the original notation we can write
or for brevity

r» = ri (gi, ...,qn\t).

(1.2.10)
When the constraints are stationary, one can choose the
generalised coordinates (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 assume this choice
ri=Ti(qi,...,qn) (i=l,...,N), (1.2.11)
where denotes the position vector of point Mi in an inertial
Cartesian
coordinate system.
Removing generalised coordinates from the 3 N equations in
eq. (9) we
arrive at 3N — n equations for the holonomic constraints. This
process is
££
feasible as deficiency of the Jacobian
Dq (1.2.12)
matrix

dqi dqn
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-st at ionary.
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 ip
between the crank and a line rotating with a constant angular
velocity u about the same axes. The Cartesian coordinates of
point M are
x = r cos (cut + ip), x = r sin (cut + ip).
The constraint equation expressing the condition of the
constant length of the crank is given by
x2 + y2 — r = 0.
2

The latter is obtained by excluding ip 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 = cut + ip. 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 generalised 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
1.3 Generalised velocities and accelerations 25

about this position is given by an infinite countable set 3 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 <ji, ...,g n. In
general, a dot denotes differentiation with respect to the time.
For example, given a function in terms of generalised
coordinates and time
f = f(qi,-,qn;t), (1.3.1)
it follows that / is given by
dt df . ^ ^ df df _ ^ df df
dt' (1.3.2)
d^qi + - + ftZqn +dqn dt ~ 2^
k=1
B^ qk +
dqk

The latter term is a partial derivative with respect to time


provided that it appears explicitly in eq. (1).
The velocity vector v* of particle Mi is known to be the time
derivative of the position vector r* of this particle. By virtue of
eq. (2) we obtain
dYj . drj
(1.3.3)
^ dq> dt

For stationary constraints r* can be taken in the form of eq.


(2.11). In this case the latter term in eq. (3) vanishes and
vector becomes a homogeneous linear form of the
generalised coordinates qs
TL r\
dvi . (1.3.4)
E T;—Qs-
s=1 ° dq

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

df_ df_ 9V£ = (fri_ dqs


dqs ’ dqs dqs'
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.
(1.3.7)

(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
#i,...,gn. The acceleration vector Wf is obtained by
differentiating the velocity vector (3) with respect to time such
that

(1.3.9)
This expression is also obtained by differentiating eq. (8)

(1.3.10)

It is easy to prove the following equalities


d dYi d dYi dvi (1.3.11)
dt dqs dqs dt dqs
Indeed, as follows from eq. (2)

Alternatively, due to eq. (3)


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 constraints (1.2) in terms of the generalised
velocities. Noticing that
Tl r\ r\ Tl c\ r\ Tl r\ r\

Ei o^ oxk
qs +
.
at ’
oxk
Vk
~
._ dyk . . dyk
dqs dt ’
qs + Zk
. _ ®Zk • i ^Zk
~ ^ dqsqs + at
s=

we obtain

( dxk , dyk dzk\


2^zJa*fc7iZ + 6ifc7iE + Cifc7iE +
S=1 dqs dqs dqsJ
fc=1 a
N [ ik~^ + bik^ + Cik^) + gi = 0 (i =
^2 . (1.3.12)
*=i ' '

Introducing the notation


N dxk dyk , dzk
=E. &ik* ~ fy/cT;- r Gfc
k=i V % —— (1.3.13)
R - j-Tr { dxk dyk dzk
at + dt
Bi — Qi + 2^ I O'ik—^r + Oik—^—
fc=i V ^
we can cast eq. (12) in the form
n
Bisqs + Bi = 0 (i = l,...,r'), (1.3.14)
S= 1

which is equivalent to the following form


n
^ ^ Bis dqs + Bidt — 0. (1.3.15)
s=1

The constraint is integrable under the following condition


dBis = dBik dBis _ dBi Q ,
’ dt ~ dqs' 1J

In the case of a stationary non-holonomic constraint, the


factors Bis do not depend on the time explicitly, and Bi — 0.
When considering particular problems, equations for non-
holonomic constraints are written in terms of generalised
velocities, that is, in the form of eq. (14).
dF1 dFx
dqi " ^Qn+m

dFm dFm
dqi ' ^Qn+m

(1.4.2)

must be zero. Provided that the Jacobian is


non-zero
dFx dFx
&Qn+1 ^Qn+m ^0 (1.4.3)
dFrn dFrn
^Qn+l
the redundant coordinates gn+i,..., gn+m may be expressed in
terms of <7i,..., qn and time t
qn+i = qn+i(qi,--,qn;t) l = (1-4.4)

Expressions for the Cartesian coordinates in terms of the


generalised ones take the form
%i — %i (Ql 5 • • • 5 Qn+m 5^)5 1
Vi Vi (Ql 5 •••> Qn+m'i t) >
=
/
(1.4.5)
Zi = Zi(qi,-,qn+m]t) )
Inserting the equations for the redundant coordinates (4) into
these expressions 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 but introducing two
additional angles (p2 and <p3 makes expressions for the
Cartesian coordinates very simple
1.4 Redundant coordinates
29

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

F\ = ai cos -f a,2 cos <p2 + &3 cos p3 —d (1.4.7)


= 0,
F2 = a\ sin -f ci2 sin <^2 s sin <£3 =
—a

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 an- 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 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.
30 1. Basic definitions

constraints whose equations are respectively


Fk (Ql j •••) Qn-\-m51) = 0 (k = 15 — 9 Tfl) ,
n-\-m
Y, aksqs + ak= 0 (k = 1, ...,r').
S=1
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 q \ , q n .
In the majority of problems some linear forms of generalised
velocities
Ql) • • •5 Qn
UJS — CLslQl T ••• T &snQn ...,77')
(1.5.1)
are favoured over the generalised velocities themselves.
Factors asi,...asn
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
^n' + l = Qn' + li =
Qn•
Quantities u;s are referred to as quasi-velocities. Examples of
these are
projections UJ\,012,^3 of 7?thecos ip + ip sin
angular velocity
T? u or projections of
the velocity sin ip, —7? sin (p + (1.5.2)
vo of a pole O on rectangular axes
ip sin 7? cos ip, Ox'y'z ' fixed relatively to the
body. ■ Cp + Ip COS 7?
The pole coordinates £O,T/O,^O relative to some fixed set of
rectangular
axes Oxyz and Euler’s angles can be taken as the generalised
coordinates
q\,..., q$ of a free rigid body. Expressions for the above quasi-
velocities are
respectively as
follows
C
Ji
U
J2
U (1.5.4)
J3
an •.. ain
&nl ••• ^nn
&n • • • b\n

bnl • • • bnn

is non-degenerate, that is, its determinant \a\ is not equal to


zero. The equations in (1) are then resolved for the
generalised velocities
Qr — br\D\ + ... + brnujn (r — 1,..., Ti).
The matrix

(1.5.9)

is the inverse of matrix a, i.e.


b — a-1, ab = ba = E, (1.5.10)
where E is the identity matrix. In other words, introducing the
Kronecker delta
6sr — f 0, s ± r, (1.5.11)
\ 1, s = r,
32 1. Basic definitions

we have
n
=
^ ^ Q'smbmr E< (1.5.12)
^'sr? m=l
m=l

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,..., „)
(1.5.13)
dqr dqk
are met. As mentioned above, failure to satisfy these
conditions does not mean that the right hand side of the
expressions for UJ is not integrable.
S

d7Ts = asidqi + ... + asndqn (s = 1, ...,n). (1.5.14)


Quantity dns is termed the differential of quasi-coordinates. As relation-
ships (1) are non-integrable, quantities TT as functions of S

coordinates do not exist. For example, let ujsdt be the projection


of the vector 0 of an infinitesimal small rotation of the rigid
body about axis Ox', however there exists no angle whose
differential is 6X> (excluding a trivial case of rotation of a body
about an axis fixed in space and the body). An introduction of
purely symbolic notation ns referred to as quasi-coordinates is
not unreasonable since this allows one to shorten equations
—JL=Us= 7rS)
(1.5.15)
with zero above (in place of a dot) indicating symbolic notation
instead of differentiation of iT with respect to time. Provided
S

7TS = ^s{qu—,qn) (s = i, ...,n),


which indicates a transition to new generalised coordinates
only.
Let ip(qi,..., qn) be a function of the generalised coordinates.
By virtue of (8) and (14) expression for its perfect differential
can ben represented
0 n r\ in
n the form
*,=x>, £
ir ir
S=1 ^ iri d(ir fri d-
r= 1 r= 1 s= 1
Ks
(1.5.16)
Here we adopt a symbolic yet logical notation
V' dri , 9ri rsWs+ dt ’ (1.5.19)

dri
dns (1.5.20)

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 r\ r\ 71
dp
dqr E dp dri _ v dvi (1.5.21)

Up until this point only linear homogeneous forms of


generalised velocities with coefficients depending upon
generalised coordinates have been taken as quasi-velocities.
A more general case is the definition of quasivelocities by
means of linear forms with free terms as?n+1
LUS — CLsiqi T ••• T G'snQn T &s,n+1 (^ — 1? •••> ^) 5 (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
^n+1 — Qn+l — 1 (&n+l,/ — ^n+1,0 •
(1.5.23)
Then we arrive at the following expressions
71+1
ajs = '^2<iskqk (s = 1, ...,n + 1)
and the above formulae hold provided that n is replaced by n +
1. Expressions inverse to (24) are as follows
71 71+1
Qr = ^ ^ brk ^^
& f c , 7 i +l ) = (v — 1, ..., Tl) ,
(1.5.25)
k=1 k=1
where
34 1. Basic definitions

1.6 Virtual displacements


Generalised coordinates q\,..., qn are functions of time which
are obtained by integration of the differential equations of
motion subject to initial conditions. 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.
dqs = qsdt. (1.6.2)
While stating the general principles of mechanics it is
expedient to introduce 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 instant 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 6qi, ...,<Sgn, are
referred to as their variations. The above set of configurations is
Qi =qi(t)+6qi,...,qZ = qn(t)+6qn, (1.6.3)
with variations 6qn 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
^2askdqk + asdt = 0 (s = 1,..., r'),

(1.6.4)
k=1
variations 6qk are related by r' conditions
n
^2ask6qk= 0,
(1.6.5)
k=1
since one must put 6t = 0 in eq. (1) for any given instant t. The
constraints of this system with n — rf degrees of freedom admit
oon~r configurations.
Let us consider a particle M given by a position vector r*.
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 displacement of particle Mi denoted by dr*.
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
without violating the constraints. The vector dr* is calculated
at a fixed time instant t retaining terms up to first order in
variation Sqthat is
n
<5r; = Vi (QI +6q1,...,qn + 6qn,t) - (q1: (1-6.7)
fc=1 °^fc

Provided that the constraints are not time-dependent, the


latter terms in expressions (6) and equations for non-
holonomic constraints (8) drop out. Differentials dr* and
(1.6.8)
k—1
belongs thus to the set of virtual displacements. A comparison
of expressions (6) and (7) as well as (4) and (5) indicates that,
in the case of nonst at ionary constraints, dr* does not belong
to this set.
What was said about the position vector holds for any
function of generalised coordinates and time ip (#i,..., gn, t). Its
differential is an increment of this function due to the motion in
(1.6.9)
...
\k=1
whereas its
variation

(1.6.10)
k=1
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
ujsdt = d7Ts = as\dqi + ... + asndqn + as?n+;Ld£.
(1.6.11)
Replacing dqk by Sq^ and setting St = 0 at fixed instant t we
obtain the following expressions for variations of quasi-
coordinates
36 1. Basic definitions

The inverse formulae are as follows


Sqr = brib'Ki + ... + bmd'Kn (r = 1,n). (1.6.13)
The expression for virtual displacement in terms of variations
in quasicoordinates is given by

(1.6.14)

where ’’the partial derivatives of 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 admitted 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 constraints at instant t + 6t.

1.7 On the commutative operations of differentiation and


variation
In the analysis that follows variations Sqs are assumed to be
differentiable functions of time. Then, by means of
differentiation of expression (6.7) for virtual displacement with
(1.7.1)

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

(1.7.2)

The difference between these expressions is


1.7 On the commutative operations of differentiation and
37 variation

FIGURE 1.2.
or, in differential form,
n
dr
ddri - 6dri = —- (d6qa - 8dqs). (1-7.4)
t^i9^
Let us consider now the path of a point Mi under the true
motion and mark the positions Mi and M[ of the point at time
instants t and t', see Fig. 1.2. Then denoting the position
vector of point Mi by we obtain Mi Ml = dri, whereas the position
vector of point M[ is + dr*. Furthermore, let M* and M'* be
positions of the point in varied system configurations at the
same time instant. These points M* and M'* are found by
means of varying the vector positions of corresponding points
of the true path, that is
MiMf = Sn,
M[M'* = 8(r;+ dn) = Sn + Sdri.
The position vector of point M* is + 8rThe variational
principles of mechanics and the methods of variational
calculus imply consideration of a sequence of varied
positions etc. as a new varied path C*.
Taking
into account that time instants t and t + dt correspond to the
positions M* and M-*, respectively, and assuming that the
position vectors of the points of the varied path are
differentiable yields
M*Ml-* = d (ri + 8ri) — dri + dSri.
Now by virtue of the evident equality
we have
dri + Sri + SdYi = Sri + drt + dSri
38 1. Basic definitions

or
d6ri — 6dri = 0 (i = 1,N). (1.7.5)
Then due to (4) we obtain

(1.7.6)

Projecting these vector equations onto axes and using


notation (2.1) for Cartesian coordinates we have

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 (s = 1,..., n). (1.7.8)
Inversely to the latter result, eq. (5) follows from eq. (8).
Formulae (8) or, in another notation
(6qs)*-6qs= 0 (s=l,...,n), (1.7.9)
expresses the commutative operations of differentiation and
variation, which
is known as the rule ”d6 = 8d”. This rule simplifies the
derivation of the-
[55].
orems in mechanics as it reduces the amount of
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. (dri)* = Svi. (1.7.10)
Let (p (#i,..., qn\ t) be some function of generalised coordinated
and time. Having repeated the derivation performed for the
position vector we obtain the result
1.8 Variations of quasi-coordinates
39

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


(Sip)9 = 6<p. (1.7.12)

1.8 Variations of quasi-coordinates


We proceed now to derive expressions for (S7rs)* — 8uos 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
n n n n
(6-KS)' = ^ (askSqk) V ask ( <%) * + ^2 Qk^2
s dask
k=1 dqr Q.ri
k=1 k=1 r= 1

n == =
n n n
huJs ^ ^ & (ttskQk) ^ ^ ttsk^Qk “ 1 “ ^ ^ Qk ^ ^ ^Qr
k=1 fc=l k=1 r=l ^

and consequently
(^7TS)# - <5^ = ^2 sk [(<%)* - <%] + E ^2
fe_1 fe lr_1
a
QrSqk'
' ( 1.8.1)
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

( <5 T T s)* - 8LO S = [(<%)* - 8qk] +


k=1
nn/ A\n nn 11

/ OCLsk OCL \ 7 \ ^ 7 r'


L L (-^r - L6fcm57rm-
sr

fc=lr=l V 't =1 m=l

To simplify the notation we introduce the Boltzmann three-index

symbols

=
(s,t,m=l,...,n).

(1.8.3)
k=1 £=1 m=l(1.8.2)
40 1. Basic definitions

and the first sum vanishes when the rule ” d6 = 6d” is utilised.
While deriving the latter result we assumed that the quasi-
velocities are homogeneous linear forms of generalised
coordinates and that the coefficients 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
(8Qn+l) — ^Qn+l — 0) <5^71+1 — 0) Mn+ 1 — 1-

The result is
n+1 n+1 n+1

(<$7TS)# - 8UJS [(<%)*-<%] + J2J2 'VtmUtfam


k=l t~l m=1
n nn n
6
= ask
[(*«*)* - &] + EE 7 ImUtfam + (1.8.4)
k—1 t= 1 m= 1 m= 1

where, by virtue of eq. (2),

£
_s _ (d sk dasr \ u u a

(1.8.5)
m— ~ 9q 9Clk 2^ [ f)n fin, I °r’n+l0fcm
ti ^ V r )
7n+l,m

/A A \ 71 /A
/ Odsk Udsr \ ( Odsk da S,n
+1 8km'
9<?
It is easy to prove that the three-index symbols in
relationships in eq. (4) may be found by formulae (2). Indeed,
we have
das d&sk \
n+1n+1

71 EE d(
brtbkn
k V lk dq r J
n n /n n /
r\ \ n
/ Od r O&sk \ 7 T , / Ud 9a
8rt8n-\-l,m
S sr s,n+l

at dqr
k=1
9a dask \ 9a (9a
8km T ^n+l + ^n+1,:
<9% “ ~ar;
s,n+l s,n+l

E
s,n+l

fc = l 57 9t
'fc,n+l^fcm

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

1.9 Some properties of three-index symbols

The values of the three-index symbols 7^ depend only on the


chosen way
of determining the quasi-velocities in terms of the generalised
7 ii • •• 7 i „

Ini • • • 'Inn

(1.9.1)

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


nn / ^ OX nn/
dasr dask \ , ,
~T, I
mt
■ »'■ > s;sV dqu OkmOrt,
uqr J
and comparison with (8.2) yields
(1.9.
2)
In particular,
7ft =0 {s,t= l,...,n).
(1.9.
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 7^ as coefficients in front of products ci 7<S7rm
appearing in the difference (£ 717)* —8UJs. There is no need to
reject the rule d8 = 8d since the concrete values of the three-
index symbols are not bound to this rule.
Expressions for three-index symbols can be transformed
into an form which differs from (8.2). Differentiating
db.km dabrt sr
dbrt
C'O’sk , E &sk ’ 1^7^7
=
~l^as
E “TJ Okm '
dqr T=i dqk
dqk
k= 1 dqr k= 1 r= 1

and substitution into (8.2)


yields (dbrt dbrrri
7 mt = 1
-------Okm----------~ (1.9.4)
\dqk dq
---------- °kt k
k= 1 r= 1

This expression for the three-index symbols is used for


obtaining equations for differences in ”second derivatives of
function cp (#i,..., qn\ t) with respect to quasi-coordinates”.
According to definition (5.17) we have
d2ip n n a ^ nn d2<p dbrk dip\
E^&E-g-EE bts^rk
dqtdq + btsdqt dqr)
d'Ksd’Kk t= 1 t= 1 r=l r
42 1. Basic definitions

Now taking into account eq. (5.21) we


obtain
r>o nn n nn ^

9 , O ip ^ dip L ^rk
fosdltk ~ ^ ^ ^ 5i ^ S S ^ Clmr
tS

and
d2ip d2ip ^ dip V^V^/^uipL dbrk L dbr
dirsdiTk dirkdns E ^ ^ \ ctyt

and we arrive at the sought-for relationship


<9V #V __ ..m dtp

dirsd7rk dirkdTTs ^ E
lks
dirn (1.9.5)
m—1
Let us consider a particular case which is important for
applications and consider m quasi-coordinates
CJS = asiqi + ... + asrnqm (s = 1,... , m),

(1.9.6)
where the coefficients asm depend on gi,..., gm, but not on
..., qn. For
the sake of symmetry we adopt
^m+;=9m+j (j = 1, . . . , U ~ 777,) .
(1.9.7)
In this case, all of the three-index symbols
7 tq (s,t,q = 1,... ,ra)

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 J5, the angles $ and x as we^ as
the angles of the wheel rotation ^I? ^2? ^3 and <£4. 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 J3, while the other four
describe the absence of velocity at points where the wheels
contact the road. Thus, the system has two degrees of
1.10 Calculation of three-index symbols for a two-axle trolley 43

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


in the following form
LO\ — 0, UJ2 — 0, CJ3 = 0, CJ4 = 0, CJ5 = 0, ( J U Q = 0, (1.10.1)

where the expressions for the quasi-velocities are as follows


uj\ = — xsin (d + x) + ycos (d + x) ?

(1.10.2)

0J2 = — x slu'd + ycosd — Id,

(1.10.3)

CJ3 = xcosd + ysmd — ad — r\<p1, (1.10.4)

CJ4 = xcosd + ysmd + ad — r\Cp2, (1.10.5)


Equations (2) and (3) are expressions for the components
of velocity at points B and A along the corresponding wheel
axes. The components
xcosd + ysmd — ad, xcosd + ysmd + ad
in eqs. (4) and (5) are equal to the velocity of the 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
- sin (if + cos ($ + x) 0 0 0 0 0 0
x)
— sin?? cos$ -l 0 0 0 0 0
cos if sin i) —a 0 —T\ 0 0 0
cos if sin$ a 0 0 —Tl 0 0
cos (if + x) sin {'d + x) —c —c 0 0 -T2 0
cos (if + x) sin + x) c c 0 0 0 -r2
cos (if + x) sin + x) 0 0 0 0 0 0
0 0 0 1 0 0 0 0

(1.10.1
1)
We obtain b = a-1 by solving equations (2)-(8) for
generalised coordi-
nates. This presents no problem as the system structure is
x = —ui sin (if + x) + u7 cos (if + x), )
(1.10.12)
y = ul cos (if + x) + u;7 sin (if + x), J

then we have
x cos $ + y sin $ = — o;i sin x + (1.10.13)
w7 cos y,
—x sin if + y cos if = UJI cos x + ^7
By virtue of eq. (3) we obtain if and then by means of eqs. (4)
and (5) we obtain generalised coordinates Cpx and Cp2- Finally,
having if we determine Cp3 and <p4 from eqs. (6), (7), (8) and
(9). The result is represented by the
- sin (1? + x) 0 0
cos (1? + x) 0 0
1 1 0
~1
y COSX 0 0 0
a 1
^51 ril

a n
&61 ril
0
C c
---- 7 COS X
r2l
0
r2l
c c
- -7 COS X
r2l r2l
0
0
0 0 cos (1? 4- x) 0

0 0
0
sin (1? + x) 0

0 0 0 0
0 0 y sinx 0 1
0
0 0 0 ^57 0

1
(1.10.14)
n 0 0
where ^67 0

1
0 677 C
^51 )•
0 T2
1
T2 \cosx~ y smx
C
0 0 r2 T2
(yx)sx+ y sinxj
00 0

/ .+ a
1 \ 1
=------- ( * 7 cos x sm bn = —
n bn — J , T\
l \ 1
= — (smx-ycos xJ ^67 = —
n For equations, of motion T\ we will need the three-index
symbolsc . >with subscripts1/ 7 and 8. Applying eq. (9.4) we obtain
00 0

=-0 - - 7 sm (l +
Xy T27 V
— y^ (^br ^ _ dbrs
778 bkr
dqk
r
a r 638 + 648(db 7 u , dbr7 <%r7
* ’O5
r=l

dx ^X
r= 1

because all the elements of the eighth column are constant


whereas those of the seventh column depend only on $ = qs
and \ — q^. Also 633 = 0, 643 =
1. Hence, we have
778 = -asi sin + x) + Os2 COS (1? + X) + 7 cos x - aS5 — (sin 0-S3

x+
l r±
a\ 1 /. a \ c c
y cos X) + as6— sm x + y cos xj ~ asi~l cos X + as8y^ cos X
and by means of matrix (11)
7^3= 1, 7y8 = sin $ sin ($ + x) + cos $ cos ($ + x) — cosx = 0,
778 = — cos $ sin (1? + x) + sin $ cos (1? + x) — y cosx + sinx + y cosx
=0
46 1. Basic definitions

etc. Next we obtain


778 "787 — 7?8 — ( s — 2,... , 8). (1.10.15)
We find the remaining three-index symbols from the difference
(6TTs)0 — 6UJs. By virtue of (14) and (9) we have

'&= -(u iC0sx-^2+^7sinx), X = ^8-


Therefore
(STTi)* — 8oJi = — cos($ + x) 6x ($ + xj — x (6$ + 6x)
sin (t? + x) by ($ + x) - y + $x)
= —b'K'j
UJ$ + y (CJI cos x - ^2 + ^7 sin x)+

<5TT8 + - (6TTI COS x — <^TT2 + £777 sin x)


UJ'j

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


are as follows
i i cos x i i J- i 1 787 ~
7l7 — “771 — j 5 727 — 772 — y> 778
o 1 9 sin x o 9 cosx
771 — —7l7 — 7l2 — “721 — j 5 727 — —772 — j 5
9 9

k k cos X 712 = -721 U Sin X n o .X


= —1—, 772 — “727 — ^ 5
U
(fc — 3,
4) ,
f = y,7?i = -7?7 = ^,7li = -7fs = 1, (fc = 5,6,7,8).
7*2 -'V
c (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.
1 X y Z

x’ an OL12 OL13
y' OL21 <T2 <T23
2
z' <^3 <^32 <^33
1

Table 1 of direction cosines


A. I. Lurie, Analytical Mechanics © Springer-
Verlag Berlin Heidelberg 2002
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's denote the unit vectors of the system
Ox'y'z'. As follows from the above table ask is the projection of i's
on the i& axis, i.e.
asfe = i,s-ifc s,k = 1,2,3,
(2.1.1)
where the dot indicates the scalar product.
i's ~ aslh + OLs2*2 + OLs3i3 — Otskh- (2.1.2)
k=1

Vectors ii, i2,13 are unit base vectors and produce an


orthogonal set. By means of the Kronecker delta one can cast it in
the form (k = m).
& km (2.1.3)
{I
Hence
33 3
*•* = ££ OLskOLtl^kl = y^^skCttk-
k=1
k=1 1=1

Since the trihedral i's is also orthogonal we


obtain
^ ^ Ot-skOt-tk — $s (2.1.4)
3
k=i

The latter yields six relations linking the nine direction


cosines of Table 1. These relations are given by
ll + 12 + 13
=
a a a
1? &11&21 + OL12 OL22 + 0,
0(13^23
a
21 + 22+ 23 ~ 1?&210131+ OL220L32 + 0,
a a
0.
(2.1.5)
a
<^23<^33
31

The vector
+ <^32 product
+a
33
ii ^31^11
x \2 is
=
defined
+ 32^12 + a
as a unit vector
&33&13

perpendicular to the plane containing ii and i 2 and points in the


direction of axis Oz, in other words ii x 12 = 13. The notation
11 x i2 = 13,12 x li -13 (2.1.6)
etc. can be shortened by introducing the Levi-Civita symbol e^im? the
value of which is given by
+ if k, l, m is an even permutation of
^klm — 1 1,2,3,
-1 if k,l,m is an odd permutation of
0 1,2,3,
an &12 £*13
&21 & 22 £*23
&31
<T32 £*33

(2.1.10)

each element is equal to the corresponding cofactor. For


instance, the elements of the first row in the determinant are
equal to the projections of the unit base vector on the axes
Oxyz, whereas the corresponding cofactors are equal to the
projections of the vector product i'2 x i3 on the axes Oxyz.
Therefore
M = ii • O2 x 13) = ii * ii = 1- (2.1.11)
The matrix

<*=\\aik\\ (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 transposed rotation
matrix (see Sec. A.4)
a-1 = a'.

(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 — ^ ^ asbris x ir — ^ ^ asbr€.sri\i.

2.1 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 Ol\t of point M and xf,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
r =xii + yi2 + zi3, r' = a/ii + y' i'2 + 2%. (2.2.1)
Let ro = OO denote the position vector of the origin O. Then
projecting an obvious geometrical identity
r = r0 + r' = r0 + x'\\ + y'i'2 + 2%, (2.2.2)
on the axes of the inertial coordinate system we arrive at the
following equations
r • ii = r0 • ii + x'i\ ■ ii + y’i'2 • ii + 2% • ii.
The scalar form of this equations is given by
x = xo + aux' + «2iy' + <*312', \ y = Vo + CX12X' +
a22y' + 0t32z', > (2.2.3)
2 = 20 + ai3x' + a23y' + a33z'. )
To derive the inverse transformation, one projects
relationship (2) on the axes of system Ox'y'z'. Multiplying both
sides of eq. (2) by ij, i'2, we obtain
x' = an (x - x0) + an (y - yo) + «i3 (z - z0), )
y' = a2i {x - x0) + a22 (y - y0) + a23 (2 - z0), > (2.2.4)
z' = a3i (x - x0) + a32 {y - yo) + a33 (2 - z0). )
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 transposition of a matrix, for example £; and x'
are row matrices.
Formula (4) takes the form
x = a(Z-£ o)-

(2.2.5)

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 enabling calculation of the nine direction cosines
ask (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 xp (0 < xp < 2 )
TT
X y z
COS ip cos ip — cos (/?
xt
sin ip sin ip cos sin ?/>+ sin ip sin ip sin d
d— sin (/7 cos cos
— sin ip ip
cos d
y' cos ip sin ip cos sin ?/>+ cos ip cos ip sin d
zf sin
d d sin ip sinipdcos
— cos cosdip cos d

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, nl5 i3 and a ”half-movable” one n, n',i's. 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

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 Oz and Oz' as in the above
example) or different names (e.g. Oy and Oxf). 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 — ni along this projection.
Then n lies in the fixed principal plane and is perpendicular to
ni.
In the moving principal plane the unit base vector n', which
is perpendicular 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
Oxfy'zf are:
1) the angle between the fixed and moving principal axes.
As an alternative one can take the angle between vector — ni
and the moving principal plane (their sum is TT/2);
2) the angle between the reference axis in the fixed
principal plane and vector n (sometimes ni or — ni); and
3) the angle between vector n (or n') and the reference axis
in the moving principal plane.
54 2. Rigid body kinematics - basic knowledge

A well-known example of Euler’s angles in astronomy uses


the following set of angles: and i determine the position of the
orbital plane whilst uo
provides the direction for a reference axis in this plane, see
Fig. 2.2. The first angle Q 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 0£rj and is the nutation angle. Angle
uo describes the spin and if the reference axis is directed in
the pericentron II then u describes the angular distance of the

2.4 Airplane angles and ship angles


It is common practice in airplane dynamics that the fixed
axes are directed as follows: O£ along the required heading in
the horizon plane, Or\ along the vertical uprising from point 0
and 0( to the right for an observer watching in the direction of
0£. 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 Oz is perpendicular to this
plane to the right for the pilot, see Fig. 2.3.
Let us place the origin 0 of the coordinate system 0£rj( at O
and view Or] and Ox as principal axes. Projecting (as indicated
in Sec. 2.3) Ox on the principal plane 0(£ we obtain vector —ni
and vector n, the latter lying in the same plane, in such a way
that i2, n and ni form an orthonormalised right-handed
trihedral, see Fig. 2.4a. Vector n' lies along the projection of i 2
2.4 Airplane angles and ship angles 55

the direction perpendicular to in the plane of the principal


axes Ox and Or] (in passing we note that vector ni also lies in
this plane).
Angles 0, d and ip referred to as the angles of yaw, pitch and roll are
generated as follows: i) 0 is the angle between 0£ and — ni
while rotating about Or?, ii) rotation through an angle p about
the axis Ox makes vector n' coincidental with Oy, and iii) the
pitch angle d is determined as the angle between vectors —ni
and i'x while rotating about—n.
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 O which is routinely the centre of inertia of the
airplane, Oy* lies in the plane of symmetry of the aircraft
perpendicular to Ox*, and Oz* 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^TyC is described by angles A, p and v which are
determined exactly in the same way as the airplane angles
0,r? and <p. Angles A and p 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 (3 and
the angle of attack a. By rotating the plane Ox*z* through an angle
(3 about axis Oy* one determines the unit base vectors m and
mi, see Fig. 2.4b, then a rotation through an angle a makes
£ c
X cos ip cos ?? sin?? — sin ip cos ??
sin^sin^— cos cos ?? cos sin (p cos ,ip+
y
(p cos ip sin ?? (p cos (p sin 'ip sin
cos (p simp+ sin (p — cos ?? ??
cos (p cos 'ip—
z
cos ip sin ?? sin (p sin (p sin ip
sin ??
X y z
X* cos a cos — cos (3 sin sin [3
(3 a
y* sin a | cos a 0
Z* — cos a sin sin (3 sin cos (3
/3 a

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 ip, ?? and p in Table 3 by A, [JL
and v. The table of direction cosines of the angles between
the velocity axes and the airplane axes is given by

Table 4 of the direction cosines of the velocity angles


X yz
X* 1 —a /3

y* a 1 0
*

0 1

Table 5 of the direction cosines


By means of this table we obtain
cos A cos p = cos %p cos d — a (sin ip sin ip — cos ip
cos ip sin $) + /3 (sin ip cos ip + cos ip
sin ip sin $),

cos p cos v — a sin d + cos d cos ip, sin p = sin d — cos $ (a cos <p
+ (3 sin p).
Because the differences between A, p, v and ip are small
values of the order of a and (3 we have
cos A cos p = cos (A — %p + ip) cos (p — $ T $)
« [cos ^ — (A — ^) sin [cos $ — (p — d) sin $]
« cos %p cos $ — — ip) sin ip cos $ — (/x — $) cos ip
sin
cos p cos z/ « cos $ cos ip — (p — $) sin $ cos ip — (v — ip) cos $ sin <^,

sin p « sin $ + (/x — $) cos $.


Comparing the corresponding equations we arrive at the
following relationships

(A — ip) sin ip cos $ + (p — $) cos ip sin $ «


(a sin <£ — (3 cos y?) sin %p — (a cos ip + (3 cos
<£>) cos ip sin
58 2. Rigid body kinematics - basic knowledge

FIGURE 2.5.

(/i — i?) cos p sin d + (y — p) cos d sin p « —a sin $,

(/j, — i?) cos « — (a cos </? + (3 sin </?) cos

$, which lead to the formulae relating the velocity and

airplane angles
p—d = — (a cos <£> + /? sin <£>), 1
\-'ip =---------- (o; sin p cos (/P) , > (2.4.2)
COS 17 I
v — p = (—a sin p + (3 cos <£>) tan i?. J

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 Oz lies in the centre plane of the ship.
They coincide with the axes 0£r]( in the equilibrium position of
the ship. Axes Or] and Oz are taken as the principal axes.
Vector — ni is obtained by projecting principal axis Oz on plane
a perpendicular to this vector defines a unit base vector n of
the nodal axis which is an intersection of the principal planes
0£( and Oxy. The angles of rotation I/J and $ about axes Or] and
n, respectively, determine the trim and the heel, whereas the
£ V C
cos ip COS (p+ — cos (p sin 'ip+
X sin 'ip sin (p sin ft sin (p cos ft sin (p cos ip sin
ft
— cos 'ip sin cp+ cos (p cos sin <p sin ip+
y
sin xp cos (p sin ft ft cos (p cos ip sin
z cos (p COS ft — sin^ cos
ft ft cos 'ip

£ c
X i1 +v ft sin (p
y -(4> + <p) | 1 ft cos (p
1
z ^sin^
—^COS^ 1

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 ft 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 0( 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 ft
and the sum of angles 'ip + p remain small for a small deviation
from the initial position. Indeed, as the angles between the
Table 2 takes the form

Table 2* of the direction cosines


y c
X 1 0
y -0 1 V
z -v 1
1 IU V c
X 1 0 0
y 0 COS Pi siiupi
z 0 — sin ipx cos (p1
1 0 0
ai = 0COS Pi sin^
0— sin cp1 COS ip1
COS 0 — sin ip 2 COS ip3 sin ip3 0
ip 2
OL2 — 0 1 0 ? ^3 = — sin cos ip3 0
sin(^2 0 COS ip2 ip3
0 0 1
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 cp1 about axis £ the table of
direction cosines of the two sets of axes Oxyz and 0£ry£ with
respect to each other take the form

If we introduce the matrix corresponding to this rotation

(2.5.1)

the matrix form of the coordinate transformation for the case £0


= 0 is as follows
x = oq£.
(2.5.2)
The matrices for the rotation about axis Or] and <p3 about 0£
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 ^ about axis 0£. This rotation brings the
trihedron £, 77, £ into coincidence with axes with x\ and y\
coinciding with axis £, the nodal axes and vector ni,
respectively. The matrix form (2) for the coordinate
transformation is given by
where a^ is matrix a% in which (p3 is replaced by if.
A further rotation d about the nodal axis, that is axis Ox 1,
puts the trihedron OxiyiZi into orientation Ox22/2^2, with axes
Ox2,Oy2 and OZ2 coinciding respectively with Ox 1, n' and Oz. Since
the rotation is performed about axis Ox 1 the rotation matrix a#
is obtained from matrix aq by replacing with d. The coordinate
transformation in matrix form is

X2 = OL'&X 1.
The third rotation makes the trihedron 0x22/2^2 coincide with
the body axes Oxyz. The rotation matrix a^ is obtained from a%
under the assumption that axis Ozs remains fixed. Thus we
have
X Oi^pX
— 2*

Now we obtain
x= a(pX2 = OL^OLfiX 1 = = at;. (2.5.4)
Thus, the rotation matrix a bringing axes0£rj( into
the final position
Oxyz is proved to be equal to the following matrix product

G, — QLtpQL'OQL'ijj, (2.5.5)
where and are matrices of the type < 23, oq and a3,
respectively.
It is worthy of note that the matrices o^, a$ and a^ 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
£ = a'x
= OL^OLfiOL^x, (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 £,77 and £.
Table 3 for the directiona =
cosines of the airplane angles is
(2.5.7)
62 2. Rigid body kinematics - basic knowledge

where 07,, a# and are matrices of the type 07,^3 and 0^2,
respectively. Table 6 for the direction cosines of the ship
angles is given by
OL — otipOtfiOtip
j (2.5.8)
where a$ and a^ are matrices of the type <^ 3,07 and ot2,
respectively.
The rotation matrix which brings the fixed axes into
coincidence with the velocity axes is given by a matrix of the
type (7)
a* = a^a^ax.

7 = aaaf3,

(2.5.10)
where aa and ap are matrices of the type <23 and 07.
or
a* = 7'a, a^a^ax =

(2.5.11)
Notice that Table 4 offers 7'. 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 (3 are small. In

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 1' of the
rotation axis of the outer gimbal (o.g.) are fitted to the
platform, whilst the inner gimbal (i.g.) 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
1 0 0
Ota — 0 cos a sin a
—sin cos
0 a a
— sin
cos P 0 (3

ap= 0 1 0
sin (3 0 cos/?
COS ip sin ip 0
— sin ip cos ip 0
0 0 1

(2.6.1)

Their elements are the direction cosines of the angles


between the axes of
the rotated body and those of the platform. Matrix a a
describes rotation
COS sin (3 sin — sin /3 cos
(3 a a
OifiOLa — 0 cos a sin a
sin/3 — cos /? cos (3 cos a
sin a
£ c
X COS (p cos ip sin (3 sin a- — cos (p sin (3 cos
cos /?
b sin (p cos a a+ sin p> sin a
— sin (p — sin (p sin /3 sin sin p> sin (3 cos
y cos (3 a+ cos p> cos a a+ cos (p sin a
z sin (3 — cos (3 sin a cos /3 cos a

This result is the table of direction cosines between two


bases, namely and Ox*y*z*. What remains is to premultiply apaa
by a^ which results is the following table

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
1 0 0 cos a.2 0 — sin a.2
0 COS /?2 sin @2 0 1 0

0 sin cos/J2 sin a.2 0 cos a.2
COS OL2 0 — sin 0L2
ot/32aa2 = sin (32 sin a2 COS (32 — sin (32 cos a.2
cos (32 sin -sin/?2 COS (32 COS
0L2 OL2

(2.6.3)

because the rotation through angle 0:2 is made about axis Oy.
The matrix multiplication yields

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/?! = cos/J2 sinai2, cos/?! sinai = sin/32, cos/^ cosaq = cos(32
cosa^.
(2.6.5)
One of these equations is a sequence of the two others.
Hence, we find
cos a2 cos ai sin a\ sin cos2 a2 cos
R
a2 R
a2 R
0 cosai — sinai
cos ai sin cos a2 sin a\ cos a2 cos ai
a2 R R R
(2

cosa2 0 — sin «2
cosa2 sinai sina2 cosai cos2 «2 sin a\
R. R. R
cos a\ sin ot<i cos c*2 sin a\ cos ai cos «2
R R R
cosy siny 0
— siny cosy 0
0 0 1

(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 0\z{ (02^). Let y denote
the angle about which the axes 0\x\ylzl must be rotated unless
they coincide with axes 02x^2^- Then

OL(32OLa2 = a^a^aai = -sin 7 cosy 0 a^aQ (2.6.11)


001
By comparing the elements of the first rows we
find
cos 0^2 _ cosy
cosa2 = —-—cosy, 0
R R ■ cos ay sin ay sin a2 + sin y cos ai,
cos y o
sin a2 = —cos2 a\ sin a2 + sin y sin
ai. R
Recalling expression (8) for R we obtain
cosy = R, siny = — sinai 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 xj; .
cos w---- — sin f tan v = — tan
ao, (2.6.13)
cost?
. tan ip
— sin f---- - — cos f tan v = tan
a\.
When yaw is absent, then
tant? = — tan 07, tan^ = — tan0^2 cost?,
(2.6.14)
which justifies the notation instrumental heel and instrumental trim for —
OL\ and — c^2, respectively. For small angles of yaw if; « —02-
Notice that tan^ = — tan/?! due to ( 6) and (14), that is the trim

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 0^rj(.
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^rf£ to Oabc and Oxyz, respectively. Then
a = Ilf, x = Cf, a = HC'x,

(2.6.15)
that is matrix IIC" makes trihedron Ofryf coincide with Oabc. Let
a denote the rotation angle of the outer gimbal of the
Cardan’s suspension relative to the base C. Let /? and 7
designate rotation angles of the inner gimbal in the outer one
and the platform in the inner gimbal, respectively. In passing
0 1 0
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 OXYZ and Oabc initially coincide.
Provided that a denotes the matrix given by Table 7 of the
direction cosines we have
a = aX = a^apdaX, x = BX, a = a^apctaB'x, (2.6.16)

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


a^apaaB'= ILC' C = Ba^a^Il. (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^T]C, only in notation. Then II = where B\ is a matrix of
notation alteration. Then
C = Ba^a'pa'^. (2.6.18)

Particularly, ifa' = /? = 7 = 0 then axes of the moving base


coincide with the base axes and
(C)a=/3=7=0 = E = BBU C = Bafaafpa^Bf. (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'pCt^ is the transpose of
Table 7 of the direction cosines
cos 7 cos /? — sin 7 cos /? sin /?
cos 7 sin /? sin a — sin 7 sin /?
sin a 7 cos a — cos /? sin a
+ sin 7 cos a + cos sin $
— cos 7 sin /? cos sin 7 sin /? cos a
+ sin 7asin a + cos 7 sin a cos /? cos a
00 -1 0 0 1
£= 01 0 , B' = 0 1 0
1 0 0 -1 0 0

(2.6.20)
Comparing C with Table 6 yields the required relationships
for the angles of yaw, trim and heel in terms of 7, a and /?. As
one can see they are rather complicated. Figure 2.10 shows
that p « —7,1? « -a and ^ « /? only for small angles.
For the second variant which is shown in Fig. 2.11 we
have

and, by virtue of (19), we obtain


— sin7 sin/? cos cos 7 sin /?
cos /?
cos a a— cos 7 sin a cos a— sin 7
C= — sin 7 sin /? sin sin a
cos f3 cos 7 cos a cos 7 sin /?
— sin 7 cos /? sin sin 7 cos a
sin a cos 7 cos /?
(2.6.21)
1 V 1 V
£ = BiE, a = —v 1 A S, n = —V 1 A
-A 1 -A 1
1 V
—v 1 A
t* -A 1
>
1 1 Z/-7 -p + f3
-p 1 d =B —v +71A—a
-0 1 fi — (3 —A To; 1

and relationship (17) yields

C = Ba^a'^a'^ (2.6.22)

Assuming that for a perfectly stabilised platform a = (3 — 7 = 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

in which matrix B is determined by the suspension design.


0 -0J3

LJ = CJ3 0 -w 1
—0J2 CJi 0

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


d^
_ = c^3i2 - ^213 = a; x ii,
dt
— UJii'3 — cosi'i = CO x (2.7.5)
dt
^3 i2,
dt
Here a; denotes the vector which is referred to as the angular
velocity vec
tor, u)i,u)2 and <^3 being the projections of vector UJ on the axes
72 2. Rigid body kinematics - basic knowledge

coordinate system Ox'y'z'. In Sec. 2.12 we will show that these


three values are transformed under a change of the
coordinate basis as components of the vector. Equation (5)
can be recast in the form
d\
- ^ = w x i ' = estqu x (i( x i^) = estq (i'twq - i'qu)t), (2.7.6)

thus
,/.d£ 3
=
1
dt ^ ] CXtm&sm = tsiq^qi

which yields expressions for the angular velocity projections


in terms of the
direction cosines and their time derivatives
di'
€stqlt CXtm&sm • (2.7.7)
m=l
For example, the full expression for is
= <^31^21 + <*32^22 + <^33<^23 = “ (<^21^31 + <^22<^32
+ <^23^33)
and so on.
By virtue of eq. (5), formula (1) for the velocity distribution
in a rigid body takes the form
v = v0 + u>x {x"i\ + y'i'2 + z'\'z),
or
V = v0 + u> x r' = v0 + wx (r - r0),

(2.7.8)
where the second term expresses time derivative of

2.7 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 Sr of a point M denoted by position
vector r = Old is normal to r. This follows from the equation
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 Sr as follows
Sr = 0 x r,
(2.8.1)
where 0 is an infinitesimal vector identical for all points of the
body. Indeed, if one considered this statement to be incorrect
6ri= 0i x ri, Sr2= 02 x r2,

one would have


r2 • Sri + ri • Sr2 = 0i • (iq x r2) + 02 • (r2 x iq) = (0i - 02) • (n x r2).
The left-hand side is a variation of the scalar product iq -r 2.
This variation is equal to zero since the length of vectors iq
and r2 as well as the angle between them do not change. For
this reason
(0i - 02) • (iq x r2) = 0,
implying that 0i — 02 = 0 since vector iq x r2 is arbitrary.
Equation (1) suggests that
Sr — |0| rsina.

With the help of Fig. 2.12 we see that the absolute value |0|
of vector 0 expresses the angle of infinitesimal rotation for
any position vector r. Therefore 0 is called the vector of infinitesimal
rotation. The definition of the vector of finite rotation will be
given in Sec. 3.1.
Provided that a rigid body has a fixed point O we have VQ =
0 and r'= r. Equation (7.8) for the absolute velocity takes the
v = UJ x r. (2.8.2)
74 2. Rigid body kinematics - basic knowledge

With this in view and setting Sr = vdt we obtain, by comparing


eqs. (1) and (2),
0 = codt.
(2.8.3)
This equation relates the above introduced vector UJ with the
vector of infinitesimal rotation 0 which explains why UJ is
referred to as the vector of angular velocity.
Let the rigid body with an immovable point be subject to
two successive infinitesimal rotations 0\ and 02. After the first
Y'— r + 0i x r,

whilst after the second rotation it becomes


Y"—r/+02 x Y' = r + 0i x r + 02 x (r + 0i x r).
It would not only be redundant but wrong to keep the small
values of the second order 02 x (0i x r) because we neglected
the values of this order in the original definitions. Thus we
obtain
r" - r = (0! + 02) x r = 0 x r,
where 0 denotes an infinitesimal rotation corresponding to the
total displacement Sr = r" — r. Therefore
0 = 01 +02 = 02 + 01?

(2.8.4)
and the infinitesimal rotations are commutative. By virtue of
relationship (3) this statement is also correct for angular

2.8 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 ip, d and
ip whereas the second position, being infinitesimally close to
the first one, is given by ip-\-8ip, $+8$ and (p + Sep. 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
0i = i^Sip, 02 = nSti, 03 = i '3S(p.
In accordance with eq. (8.4) the resultant infinitesimal
rotation is just a geometrical sum of the components
0 = isSip + nSd + ig Sip. (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 velocity vector
u = 131P + nO + i'3ip.

(2.9.2)
Its projections on the axes that are fixed in the body are
oil = ui • i[ = ip sin 0 sin ip + 0 cos ip, "j
0J2 = u; • i'2 = ip sin 0 cos ip — 0 sin ip, >
(2.9.3)
£1 = u; • ii = ^ cos 'ip + Cp sin 0 sin 'ip,
£2 = u • i2 = fisimp — ip sinO cos'ip, >

(2.9.4)
£3 = u • i3 = ip + Cp cos •&. )
Analogous equations can be constructed for projections of
the infinitesimal rotation vector. To this end, it is sufficient to
replace the time derivatives ip, 0 and Cp in eqs. (3) and (4) by
variations Sip, 80 and
61 = respectively.
Sip,sin 0 sin ipS'ip Then
+ one obtains the
cos ipSO, (2.9.5)
62 — sin 0 cos ipS'ip —
sin ipSO,
63 = cos OS'ip + Sip,
1 = sin 0 sin xpSip + cos
'ipSO, (2.9.6)
2 = — sin 0 cos ipSip + sin
We notice at this 'ipSO,
point that the angular velocities of the ” half-
fixed” trihedron n, rq, i3 and the ” half-moving” trihedron n, n^, i'3
are equal to 1/113 and ipi3 + nO, respectively. Due to formulae
(7.5) we find n
i = -^n, ^=0 (2.9.7)
n — '1/113 x n = 'i/ini,

and
n = ipni = ip (n' cos 0
— i'3 sin 0), "
n' = — Uncos'# + ^3, (2.9.8)
^ = V>nsintf-tfn'.

Expressions for the variations (5n,(5ni, (5n' and 8i's, which are
infinitesimal displacements of ends of the corresponding
vectors n, rq, n' and i'3 subject to infinitesimal rotation 0, are
obtained by replacing 'ip, 0 and Cp in eqs. (7) and (8) with
variations Sip, SO and Sip,
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 UJ of the angular velocity
S

vector and projections vos of the velocity vector on the axes


fixed in the body as quasi-velocities. We keep the notation
U\,U2 and oos for the first group of quasi-velocities and denote
the projections vos by ^4,^5 and COQ. This ensures that there are
\ • 62 • 5 = 90 three-index symbols to be calculated.
1. Quasi-velocities of the first group are related to the
generalised velocities by means of formulae (9.3). The
ift = ——- (ui sin p -\-LO2 cos p),
. smv
(2.10.1)
d = u\ cos p — (J2 sin p,
p — — (ui sin p + 002 cos p) cot d +
Variations of the quasi-coordinates Sir 1, Sir 2 and Sirs are
equal to projections of the infinitesimal rotation vector on the
axes fixed in the body. Their expressions in terms of
variations Sift, 6$ and Sift of the generalised coordinates are
defined by relationships (9.5). We obtain the inverse ex-
pressions from eq. (1) by replacing ift,d,p and oos with Sift,Sd,Sift
and 6S, respectively.
We start with the following equation
{Sir 1 )* = (sin d sin pSift + cos pSd)* , Suo\ = S (ift sin d sin p + d cos
p^ .
Accounting for the ’’rule dS = Sdv for calculating the difference
between the above quantities we obtain
{Siri)# — SUJI = (dSift — iftSd^j cosdsinp +
{pSift — iftSp) sindcosp + (dSp — pSd^j sinp.

In this equation it is essential that the generalised velocities


and variations 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
{Sir 1 )* - Sou 1, {Sir2)m - SUJ2, {Sir3)# - S003,
we arrive at the formulae
(^7Ti) — Suj\ = u)^Sir2 —
LO2SIT3, {Sir2) — SUJ 2 — (2.10.2)
oJiSir^ — UJ^SlTi, {Sir 3) —
S003 = UJ2SIT\ — uj\Sir2.
2.10 Calculation of three-index symbols 77

By virtue of (8.3) we have


732 = -723 = 1, 7?3 = -731 = 1, 721 = “712 = 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
CJ4 = an±o + auyo + ai3z0
etc. which are obtained by projecting eq. (7.2) for the velocity
vector of pole O on the axes fixed in the body. This calculation
would be very cumbersome. It can be simplified by using the
relationships
^3+s = V • i'SJ <S7r3+s = <5r0 • ig,
0

(2.10.4)
as well as the formulae for differentiating (7.5) the unit vectors
6i'a = 0 x i's. (2.10.5)
In eq. (4) <5ro 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 < 57rs. Then we have

(<57r3+s)* = (£r0)# • i’s + <5r0 • (w x i'J, 6M3+3 = 6v0 • i's + v0 • (0 x i^).


Taking into account relationship (1.7.10) we find
(<S7T3+s)# - <Su>3+s = (<5r0 x w - v 0x^) - i's. (2.10.6)
By setting s = 1, 2,3 we obtain the equations
(81X4)* — 80J4 = CJ3S7T5 —OO281X5 ~(<^5<$7r3 —
OO581X2)
1
(^7T5)# - 8LU$ — LO181X5 -u)38iX4 —{LO581X 1- OO48IX3),
>(2.10.7)
Let us recall that 004,005,005 and 81x4,81x5,81x5 denote the
projections of vectors v0 and <5r0 on the axes fixed in the body,
respectively. The following twelve values are non-trivial
735 = -753 = 762 = - 7 2 6 = 1, >

7l 6 — — 7 6 1 =
1? 743 =
—734 =
1?
> (2.10.8)
= =
724 ~ —742 1? 751 ~ —7l5 1- ,

3. Provided that the projections 00s of the angular velocity


are taken as quasi-velocities and using eq. ( 9.4) and the
inverse equations in the form
0 = aq cos ^ + Co2 sin
<P = -—a (&i sin ^-002 cos (2.10.9)
^),
sin v
— 00 3 — cot d {oo\ sin — 002
cos ^ji)
78 2. Rigid body kinematics - basic knowledge

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


relationships
(<57TI)# — 8uo i = u2Sn3 — u)367r2> 1
{8^2)* — 8u>2 — UJ387T1 — d}i^7T3, >
(2.10.10)
(87Ts)m — 8uJ3 = OJ187T2 — OJ287T1. J
The three-index symbols are as follows
=—
723 732 =
1? 731 = —
7l3 =
1? 7l2 =
“721 — 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',^ introduced in Sec.
u/ = + ?/d3 = #n + i/j sin^n' + -0 costfi^. (2.10.12)
The angular velocity vector of the body is given by
u: — {jJ +

(2.10.13)
Let us take projections of u) on the axes of the ” half-moving”
trihedron as quasi-velocities
uj\ = uj2 sin f?,
=
0J3 = -0 cos$ + (p.

(2.10.14)
The variations of the quasi-coordinates are then
87Ti = <5$, ^712 = sinf?^, ^713 —
The three-index symbols with superscript 1 are equal to zero
since d is a ”true” coordinate.
5. Let us now take projections of the velocity vector VQ of a
pole O on axes of the ”half-moving” trihedron n, rq, i 3
00 4 = v0 n = XQ 'ip T yo sin "j
• COS -0,

3 = vo • ni = — xo sin^ + yo cos^, >


(2.10.17)
oo6 = z0. I
The latter expression is integrable and the three-index
symbols with superscript 6 are equal to zero. As the quasi-
velocity 002 is given by eq. (14) we have
(8714)* — 8LU4 = — (i/sSxo — XQsin^ + (i>8yo — yo^i^j cos-0

= I/JSTTS — ujr>8'ij) = {0028113 ~ ^5^2)


sin#
2.10 Calculation of three-index symbols 79

and we obtain
4 4 1 1
sin'#
725 “ ~752 —
742 — “724 ~
sin$’ (2.10.18)

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 i 3 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
vp = v0 + u x (rp - r0) = v0 - u x i3a = v0 — au2h + a&ih,
as rp — ro = — i3a, a being the sphere radius. As above, UJ are S

the projections of the angular velocity vector on the fixed


axes.
Equations for the non-holonomic constraints are
Xo — auj2 — 0, 2/o + = 0.
The system has three degrees of freedom. Let us take
(DI,U>2,£>3 as quasivelocities and introduce
UJ4 = Xo — au2 = xo — a — (p sin'd cos
sin 0 (2.10.19)
+ Cp sin'd sin
£>5 = 2/0 + a£> 1 = yo + ($ 0
acos

Due to equations for the non-holonomic constraints, £>4 =


0,0)5 = 0. Then we have
(^4)* — <5ct>4 = —a [(6^2)* — Su2\ , (Sir5)* — 8UJ§ = a [(STTI)9 — 8Q1]
and, by virtue of eq. (10), we obtain
713 = -731 = 723 = -732 = a-
(2.10.20)
The remaining three-index symbols are determined by eq.
(10). The general 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 XQ, 2/O5 $
and ip as in the case of the sphere. They are shown in Fig.
80 2. Rigid body kinematics - basic knowledge

FIGURE 2.13.

coordinate zo of the centre of the ring is expressed in terms of


angle i9 by means of the following relationship
z = asini?. (2.10.21)
In the case being considered
rp — r0 = an', (2.10.22)
where a is the ring radius and n' denotes the unit vector of the
half-moving trihedron n, n', i'3 directed upward along the
steepest line of the ring plane. The velocity of the contact
point P is given by
vp = vo + x (rp — r0) = Vo — au> x n'
= v0 — a (no;i + nfoj2 + 13^3) 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 = v0 — a\3uji + ano;3 = 0.
(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

CJ4 H-ao;3 = 0, ou$ + sin?? = 0.

(2.10.24)
Projecting on axis Oz yields the following integrable
CJ4 = 0U4 4- ao;3, cjg = cue 4- acu\ sin??, (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)
(<57r!)* — So4 = ^ - (CJ2<5TT5 — ^5^2) + a (02S7V1 ~ ^1^2) •
smv
Replacing here STT^ and 05 by their values from (26) we obtain
(STT4)* - S04 = ^7^ (<^2^5 - ^5^2) •
(2.10.27)
(«*5)# - ^5 (CJ2^7T4 — (JU467V2)
sin$
or replacing Sir4 and 04 by their values from (18) we have

(671-5)* - = -T—5 (<^26774 - 0746772) + (W26773 - -^-7

W36772) .
Hence, the three-index symbols are as follows
1 (2.10.28
4 4 sin$’
725 — “752 — “7 )

742 ~ “724 “7

sin$’
723 — 732 — “7

sin$ (2.10.29

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 + Co x r' + o x (a; x r').
(2.11.1)
Here wo = Vo designates accelerations of the pole and r' = r — 1*0
is the position vector with respect to the pole. The time
derivative of the angular velocity vector Co is referred as to the
e = u = J2 (w,i's )* •
S =1
82 2. Rigid body kinematics - basic knowledge

Applying the formulae for differentiating unit base vectors


(7.5) we obtain
3 3 3 3

£= + x i's = +u x (2.11.2)
s=1 s=1 s=1 s=1
Thus projections £s 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 es 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 £s of vector e on the axes fixed in space are also
equal to the time derivative of projections Cus of vector u on
these axes.
We consider now a trihedron of moving axes which are not
fixed in the body and denote the unit vectors of this trihedron
by i*. Let u/ and a;* = a;- 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 = Co — cj*i* -bo/ x u = cj*i* — (QJ — a/) x u. (2.11.3)


s=1 s= 1

Here u — a/ is interpreted as the angular velocity of the body


£ = On T sin n' + (ip cos $ + (pj i3 —
i'3<p x n$ + n'ip sin $ + i3 (ip cos $ + ^ j — (@ + ^ sinn +
^ (2.11.
4)
In formula
(ip sin $ +(2) for the
ipi) cos $ — n' acceleration
+ (ip cos $ —inipi}the
sin rigid
$ + ip'jbody the
second term £ x r' is referred to as the rotational acceleration whilst
the third term u x (a; x r') is named the centripetal acceleration. The
absolute value for the centripetal acceleration is
\CJ x (CJ x r')| = uj2h, (2.11.5)
where h is the distance between the rotation axis passing
through O and the point under consideration. This acceleration
is directed toward the rotation axis.
Expression for the centripetal acceleration can also be cast
in the following form
u x (u x r') = u) • r') — UPY' — (CJCJ — Eu; • cv) • r'.

Here UJUJ is a dyadic and E is the unity tensor. The basics of


the dyadic calculus is briefly outlined in Sec. 4.3.
(in = 0 CO 12 — —tO Cd 13 = OO 2
Co = Cd 21 = 00 3Co 22 = o C023 — —0O\
3 31 =
Cd CO32 = OO1 C033 = 0
—002

(2.12.6)
84 2. Rigid body kinematics - basic knowledge

is given by the three following values


LUi= —^23 = ^32, ^2 = —^31 = ^13? ^3 = — ^12 =
^21-
Using the Levi-Civita symbols the latter can be cast in a short
form
0Jr = -erstCost (r = 1 , 2 , 3 ) . (2.12.7)
It will be shown below that oor can be treated as components
of the angular velocity vector uo. Equation (4) takes the form
v — vo = oox (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
a' ( v - v o )= £ - £ o = a'ua (£ - £ 0 ) , (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
00 = Oi OOOi — OiOiOiOi = Qt! QL (2.12.10)
determines the column-matrix dor of projections of vector oo on
axes fixed in space. Analogous to (7) we have
00 T — 6rsfLV sf. (2.12.11)

It is necessary to prove that two trihedrons oor and oos


transform like components of a vector under a coordinate
transformation, i.e.
— ^ ^ Oisr^r
• r= 1
Due to (10) we have
00 = (2.12.12)

OLOOOi

^ ^ &skVkl
3 3
. 3 3

oo = a EE OLskOLtlUkl EE ^sk^tl^qklV q

k=1 1=1 k=1 k=l 1=1


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

oo ^str E OLj'qOO q
q= 1
— sin p — COS p 0 0 -1 0
cos p — sin(^ 0 (p, ava'v = ip 1 0 0
0 0 0 0 0 0
0 -1 0 0 0 0
d) = 1 0 0 + 0 0 -1
Cp fiOLpp
0 0 0 0 1 0
0 sin'd cos p —'d
^
— ^p + ^ cos
sin p
—^ sin $ sin p —'d
Cp + ^ cos $ 0
cos p
—^ sin $ cos p> ^ sin $ sin p 0
+i? sin p +1 ? cos p

etc. Applying eq. (13) we obtain

OLu

0 - 1 0
-\r 'IpOttpQLft
1 0 0 OL#OL
1 0 0

(2.12.14)
which, due to eq. (7), is equivalent to formulae (9.3). Matrix u is
calculated 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
£-£o=£(£-£o)- (2.12.15)
Differentiating this equality with respect to time
yields

l-£o = £(£-lo)+^(£-£o) = ( ^ 2 + i
) (£-&>)>
(2.12.16)
0 — UJ2 2 2
—UJ ~b UJ
2
UJ1 UJ2 UJ1 UJ3
UJS
UJS 0 - UJl = UJ2UJ1 —UJ^ + UJ2&3
— UJl 0 UJ3 UJ1 UJ3 UJ2 —UJ + UJ3
UJ2

— —EUJ UJ -f- UJUJ ,


where UJ' is the row-matrix || UJI , a>2» £ 3 ||- This relationship
follows immediately from eq. (A.4.7). Therefore,
I - So = (^' - Eu/u + !)(£- £0) •
(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 - W Q 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 ^ = a {UJUJ' — EUJ'UJ + e) a'x. (2.12.19)

But auj = uj,uj'a' = UJ' etc. and hence due to eqs. (10) and (5)
OiEOi — a (a;)* a' = a (a'uba)* a' = aa'uj + (cl;)* + ubaa'
= UJ2 + (UJ)* + UJ (UJ)' = (UJ)* = £,
since UJ = — (UJ)' is a skew-symmetric matrix. Hence
(cl;)* = e = OiEOi — Oi (cl;)* a'.

(2.12.20)
This expression, being analogous to eq. (17), expresses the
skew-symmetric matrix of angular acceleration e in terms of
projections of the angular acceleration vector e on axes fixed
w — wo = (UJUJ' — EUJ'UJ + e) x = {UJ2 + e) 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
e = (cu)* = (OLOLY = aa' + aa (2.12.22)
and
(co)2 = —Cu'Cu = —aa'aa' = —aa . (2.12.23)
Thus we obtain
u)2 + e = aa', w — wo = a a x , (2.12.24)
where w — W denotes the column-matrix of projections of vector
Q

w — Wo on the axes fixed in the body. In view of (2) we find

£-£0 = a'x.

(2.12.25)

2.12 Differentiation of vector in a moving coordinate system


We consider a vector a whose projections on the trihedron of
unit vectors i'l5 ig, ig, rotating with angular velocity co, are
ai,a2,c&3, respectively. In this case
a = dii'i + ^212 T ^3^3*
(2.13.1)
While differentiating this vector we should account for
formulae (7.5) for derivatives of the unit vectors. Thus we
obtain
a = diij + x + CL2U x i '2 + x i'3
— CLll'i + CL2^2 T ^3^3 T ^ X ( t t i q + ^2^2 ^3*3) •
Let a denote a vector whose projections on the trihedron
are as,
that is
a= ci\ 1^ T U2I2 T &3I3.
(2.13.2)
Equation (2) then takes the form
a =a +c0 x a.
(2.13.3)
Vector a can be referred to as the relative or local derivative of a, the
term u x a takes into account rotation of the trihedron , V2, i3 on
which the vector is projected.
88 2. Rigid body kinematics - basic knowledge

2.13 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

,x'N,y'N,z’N) =0 ( 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
x'i = x ' i( qi , . . . ,<? „) , 1
y ' i = y ' i( qi , - - - , q n ), > ( 2. 14. 2)
A = z ' i ( q u . . . ,q n ) , )
or, in vector form,
A = A(QU--- ,«»*)• (2.14.3)
Notice that these relationships do not contain time explicitly.
In what follows, subscript i is omitted.
The position of axes Ox'y'z' is prescribed by coordinates #o,
2/o> 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 ) , 2/0 = yo ( t) , ZQ = ZQ (t) ,1 4

* ( > = ip ( t) , = < p = < p( t ) , J ’ ’

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
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 acceleration 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 u and the angular
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 determined 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
n (3), with respect to time 2
f)vr n /
0 nn
d r' . .
vr w O r\ QkQs- (2.14.5)
” oqkOqs
S=I s= 1 ” ,s=l k=l

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

r, wa r = va (2.14.6)
Alternatively, differentiating eq. (2.2)
r = r0 + r' = r0 + x'i^ + y'i2 + 2%,

(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
which can be also obtained from eq. (13.3) by replacing a with
r'. Noticing that
U ', dz'., A dr' .
*r _ \ ^ / dx’
;/
UJ
1 dyy ;/ 1 uz
(2.14.9)
-L( + -HZ'*
dqs dq +saZ 3 (Is
S=1

we arrive at the theorem of addition of velocities


vo=v0+uxr' + vr=ve+ vr, (2.14.10)
90 2. Rigid body kinematics - basic knowledge

which expresses the absolute velocity as the geometric sum


of the translational and relative velocities.
Since vectors TQ and do not depend upon the generalised
coordinates and r' is independent of the generalised
velocities, it follows from eqs. (4), (8) and (10) that
dr _ dr^
_ ch_ &Va , _ 1 (2.14.11)
dqs dqs dqs dqs
Differentiating eq. (10) with respect to time leads to the
In order to prove the theorem we
theorem of addition of accelerations.
should repeat the derivation of formula for acceleration in the
rigid body (13.3) for vectors r' and vr. We obtain

wtt = w0 + e x r' + cox (CJ x r') + cox r + vr x vr.

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


vr= w, (2.14.12)
r?

Therefore
wtt = w0 + e x r' + UJX (co x r') + wr + 2CJ x vr,

(2.14.13)
wa = W e + Wr + W Cor- (2.14.14)

Absolute acceleration is equal to the geometric sum of the


translational acceleration we, the relative acceleration wr and
the Coriolis
acceleration wcor- The latter is given by the formula

WCor = 2w x vr.

(2.14.15)
We proceed now to the formula which is important for the
forthcoming. Let a denote a vector depending on generalised
coordinates </i, ...,gn, then
n
a =a x a = ——qs + UJ X a.
In particular,
2.15 Absolute acceleration of point moving over the rotating
earth 91
2.14 Absolute acceleration of point moving over the rotating
earth
Let VN , vo and denote components of the velocity vector of a
material 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 tangent
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 4> and A, respectively, and h denotes the
height above the origin O of the axes system Oxyz.
Denoting the radius of the earth as R we have

The components of the velocity vector v of the geocentric


system Oxyz along the axes of this system are equal to
•R
v\ = R$ = 7 VNi
R+h
V2 = —R ^A + cos 4> =
R (2.15.2)
R + vo + RU cos
V3 = o, h

whereas the components of the angular velocity are given by

cvi = (u°) =vo+ + U cos4>,


cos 4> R
UJ‘2 vN h (2.15.3)
R + h’
(A + [/) sin$ = vo
= U cos4> ) tan<E>,
R+h
U3 =
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
Vj.\ — 0, vr2 0, Vr 3 = h V .
= h (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,
92 2. Rigid body kinematics - basic knowledge

the following expressions for its projections on these axes


Val = Vi + U02h = VN,
va2 = v2 — h = —vo — (R 4- h) U cos$, (2.15.5)
Va3 = h.
We proceed now to calculation of acceleration. Acceleration w
of the origin O of the axes Oxyz is given by
R R + RU cos $ ) if2
w=v= VN h ~ VQ
_R + h R -\- h
Applying eq. (7.5)
yields R
w VN + vo + RU cos $ ) UJ i'i
3 +
R -\- h
R
v + RUcos*\ 3
R+h
R+h 0

R R
-VN(V 2 + vo 4- RU cos<I> )
R 4" h \ R 4"
h
The projections
R areRgiven by VQ
wi= - VN ~ j V N V h 4- R 4- U cos $ ] tan $
R + h " ( R + hf R 4"
R. R ^ V N h. ^ RVNVO ^
w2 = -pr~Tvo + “I——voVh + 2RU ^ , T sin $ + ----------------— tan
R + h ~ {R + hf R+h (R + hf
2
R o ^ 4- U cos $
w3 R
{R + h)2 N
U
R( v+0 h
(2.15.6)
To construct an expression for the translational acceleration
it is also necessary to calculate the following vector
e x r' 4- w x (CJ x r') = e x r' 4- UJUJ • r' — u>2r',
where r' = hif3. Projections of this vector on the axes are
(e2 + W1U3) h, {-£1 4- UJ2LU3) h, - (u;2 4- a;|) 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.

vo VoVh v
£1 R + h {R + hf - 7rk'***'
VN vNvh
(2.15.7)
£2 =
R + h (R + h) 2 •
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

W Cor = 2 w x i'3vh = 2i[u)2h - 2i'2u>1vh.


Now we can determine the absolute acceleration. Its
projections are
=w
w
a\ \ + (£2 + W1W3) h + 2u)2Vh,
Wa 2 = W 2 + (Si + U> U>z)h - 2u}\V ,
2 h

Wa3 = Wz~ {u)\ + U>1) h + Vh-


By means of the above formulae we obtain
. VNVh v“n tan <3? „ _T . ^ , '
wai =vN+ +0 ■■■ + 2v o U sin $ + (R + h) U2 cos
$ sin <f>,
K + n H +n
W a 2 - - V o ~ ~s~r + ^TT tan ^ + 217 (vN sin $ - vh cos $), >
H tl rt + h
wa3 = Vh — ^ p ~ 2f7vo cos $ — (R + h) U2 cos2 R h
(2.15.8)
Another form of these equalities is given in [69].

2.15 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 O of the body. We introduce the
Gaussian coordinates q1 and q2 on surface 5, then the position
vector p with the origin at the pole O is considered as a
function of q1 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 q1 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

Q1 (t) - Q2 ( t ) , x ( t ), y ( t )

prescribe curve L' which is a locus of the contact point M on


94 2. Rigid body kinematics - basic knowledge

2.14. Curves V 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 p1 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 LO and applying relationship
(13.3), we have
m =m -\~uj x m = 0,

(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)
m= maqa = -ba/3p0qa, (2.16.2)
where p@ are the contravariant basis vectors and bap are
coefficients of the second quadratic form of the surface.
which yields
—bapqam x p@ + u> — flm = 0.

But
p0 xm =—p0x (pt x p2) = -T= ~ Pia() ■
(2.16.3)
V\<i\v
V|o| \/\a\ y
2.16 Body rolling on a fixed plane 95

Here aap denote the coefficients of the first quadratic form, i.e.
the covariant components of the metric tensor of the surface,
af stands for its mixed components and \a\ denotes the
discriminant of the first quadratic form. Now we obtain
u) =f2m+ (baip2 ~ &a2Pl)<f*, (2.16.4)

which gives an expression for vector UJ in terms of the


quantities determined by the surface geometry and f 1.
Projecting the latter equation on the axes fixed in the body
we have
, dx' dx'\ , ( dxf
uj\ = Qrrii + 1 b „ 1— - b 1 2 — ) q + ( k i ^ s
7 d x ' . ,2
b22
WU
i'
7R - dq2 dy' dq2 dy' b2
UJ2 = ftrri2 + i w dq
+
\tb 2 O
^- ^)r
<U dq
1
dz'
! _b 2 2 dz'
^3 = + d q - ~ b l 2 / dWq K+ 621
75^2 “ ^ T Id q
1
dz'
dq 2
(2.16.5)
In order to understand the geometrical meaning of
parameter f t we notice that vector r can be represented in two
ways. First it is the derivative of the unit tangent vector to the
curve L and for this reason
r = i 9m x r. (2.16.6)
Alternatively,
• * 7 1 *<* .
r =r +Ct? x T = mbaQq —----------------b pak a + u) x r. (2.16.7)
da

Here relationships (B.8.3) and (B.8.6) are used, k * a denotes


contravariant components of the vector of geodesic curvature
and d a is the arc element of curve V . Substituting expression
(B.8.2) for r we obtain
-L
UJ X T —

dq@
Q,m x rH—j= (balp2 - &a Pi)<f* X — p/3
2

da '
= flm xr —

As one may expect, the term corresponding to the vector of


the normal curvature in eq. (7) vanishes. Inserting the latter
expression for a; x r into
96 2. Rigid body kinematics - basic knowledge

eq. (7) and comparing the result with eq. ( 6) we arrive at the
following equation
d = £2 + pa • (m x r) k* a. (2.16.8)
In addition we have
Pa ■ (m x T) = “4=i IP* X
(PI
x
Pa)] • P/3^
A/H
dqP
1 (^l/3^a2 — (^2/3^al)
/r-r
“7“
vlal da

and
— r.—r (^1/3^Q;2 ^2(3^cx.l) ^
pa • (m x r) vM
= A/R ■

Thus we have
')■ (2.16.9)
tf = Q + y/\a\ (q1 k* — q2k*

It follows from this equation that for given q1 (t), q2 (t) and ft (t)
angle d is determined by a quadrature, and then x and y are
determined by quadratures from the following relationships
dx = cos dda, dy = sin dda (2.16.10)

since in the case of rolling without slipping the arc element da


is the same for curves L and V. This problem is solved in Sec.
8.3.
Let v denote the velocity vector of the pole O 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
f f
X i = v i +o o 2 z -u 3 y = 0,
"I
X2 = v
2+ = 0, >
u3x' - (2.16.11)
X3 = ^3 +^12/' - u 2 x f = 0. J
Next we introduce the notation
X5 = 91, Xe = 92, (2.16.13)
where uo\, a>2, ^3 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 <$7rs. TO shorten the equations we introduce


vectors % and <$7T whose projection on the axes fixed in the
body are xs and <$7rs (S = 1, 2, 3). Then
X = v + wxp, <$7r = br + 0 x p,

(2.16.14)
6 —buj = 0 x u (<$r)* — 8v = 8r x u — v x 0. (2.16.15)
Here an asterisk designates the time derivative of a vector
relative to axes
*
Ox'y'z' fixed in the body. This implies that projections of vectors
6 , (bn)* etc. are equal to the time derivative of the projections of
these vectors on the above axes. Thus we obtain
(bn)*— 8x = br x u; - v x 0+ (0 x u) xp + 0x p —u> x bp. (2.16.16)
Replacing now <$r and v by their values due to eq. (14) and
taking into account that p and bp are equal to paqa and pabqa,
respectively, we arrive at the following expression
(<$7r)* - bx = bn x u - x
x
#+ (0qa u>bqa) xpa

(2.16.17)
where the following terms vanish
— (0 x p) xw+ (CJ
x p) x 0+ (0 x u>) xp = 0.
What remains is to replace u> and 0 on the right hand side of
eq. (17) by their expressions due to eq. (4). The result is as
follows
mx pa vlaAl—r ( ctlP2 ct2Pl) — k
a a
a (2.16.19)

we obtain
(b n ) * - 6 x = ( X 4 ^ - x N x m + (2.16.20)
(q b7r - X$q ) x la + (q bn4 -
a a
K- a

The right hand side contains only quasi-velocities and the


corresponding variations of quasi-coordinates. By projecting
vector relationship (20) on axis Ox' we obtain the three-index
symbols Yap which are coefficients of the products where
q = X5 and q = X6 due to eq. (13). The tables
1 2

of the three-index symbols having superscripts 1,2 and 3 are


a \0 1 2 3 4 5 6
10 0 0 0 0
2 0 -777,3 il3 ^23

3 777,2 il2 h 2
4 -fen ~k2i

5 0
a\P 1 2 3 4 5 6
10 0 777,3 *13 I23
2 0 0 0 0
3 —777,1 — ill —^21
4 -&12 ~k22
5 0
a\/? 1 2 3 4 5 6
10 0 -777,2 — il2 Table
—I22 of 7* ^
2 0 777,1 ill hi
3 0 0 0
4 — *13 ~k23
5 0

Table of 7^3

Table of 7^
Here kas and las are interpreted as projections of vectors on
axes Ox'y'z' for example
1 / dx' dx' ^ <9#' dx'
=
7Pirwaiw *2 a21^-a22 —
1( dxf dx' \ 1
,b21
^2 df~ 22W)

The difference (^^4)* — 6x4 can be written as


follows (<57T4)* 6x4 = (m • 0)* - 8
(m • ui) = m •
2.16 Body rolling on a fixed plane 99
as m = 0 and <5m = 0. Substituting expression (10.10) for
difference 6—6UJ and expressions (18) for 0 and UJ into the
latter equation we obtain
= m • (la x 1/9) qaSq0.
(6TT4)* -6X4 (2.16.21)
Using projections of vectors la and 1^ and recalling the
definition of m we find that
m • (Q x (balb/32 ~
la) = —b7a==2b/3l)
1
,
vH
hence

(8n4)* - 8x4 = -6'1—Tf-r1- (<?W - ifbq1) ■

(2.16.22) v\a\
Due to eq. (20) the non-zero three-index symbols having
superscript 4 are only
4 4 b\ 1622 — b\2
756 = -765 (2.16.23)
=- -VW\
All three-index symbols having superscripts 5 and 6 are
identically equal to zero since expressions x 5 = q1 and x6 = q2
are integrable.
We determine the angular acceleration vector by
differentiating eq. (18) for the angular velocity vector with
+ 1 aq + maQqa + 1 apq q^-
a a
e = UJ

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

ma- bip0- ba0pP, °e-(™dq0^


1 d ba i d ba 2 .
P2 rr-r Pi ”7== I +

{6 20m} + &2/3m
) _ ba2
({ 1/3 } p
~<
+

y/W\ W

or
m
(baib2p — ba2bif3) +

d b, a2 ^ (bal | ^ } ba2 | ^ |
Pi + +
W V W VW
d b.
+ 6al
P2 ^^ ^( { w } ba2
{ i}
100 2. Rigid body kinematics - basic knowledge

FIGURE 2.15.
Noticing that
(baib2(3 ~ ba2b10) =0
we obtain

bltflqa + qaqP ^ ^a2 + %?“ + 1


u= $lm — pi ba 2
dq0 y/\a\ ^/\a\ y/\a\ U)
- 6 “ i { 2 /3} ) « V + P 2 -b2attqa + qa® bai ..Q

vWH — 7=q
VVH
• *1 *2
{ 20 } ~ ^ { }) ^ = mO + Pi cj +p2 ^ •
+ V\<*\ 'Oil
(2.16.24)

As an example we consider the case of a rigid body


bounded by a surface of revolution. Axis Oz' is directed along
the axis of symmetry of the surface, then
x' = v (s) cos tp, y' = v (s) sin p, z' = z' (s),

see Fig. 2.15.


The Gaussian coordinates are the arc s measured along
the meridian arc from the point of intersection of the meridian
and axis Oz' and the azimuthal angle p. 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
Ox' Oy' Oz'

Pl cos a cos a sin — sina


cos <£? <£?
P2 —vsmp v cos p 0
m sin a cos sin a sin p cos a
^

Then we find
an Pi ' Pi ~ a
12 = Pi • P2 = <^22 — P2 ' P2 —
U<2
1
N/R
. 9m da

^12 9m ^ . 9m .
• p2 = 0, ^22 = * p 2 = ~^sma,

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


Christoffel symbols of the second kind for the surface of
revolution are COsa
’ { 12 } = { 221 } cos
a
v
whereas the components of the vector of the geodesic
curvature for the curve s = s(a) and ip = ip (a) are given by
d2s dip\2
— v cos a, k*2 = d (p
2
fe*1 da2 dip ds cos a
dcr
da2 da da v
By virtue of eq. (9) we obtain
d = SI + v ((2.16.25)
\da da )
Formulae for projections of the angular velocity vector on the
axes Ox'y'z' take the form

UJ 1 = (12 + <p cos a) sin a cos cp +


ks sin p, (2.16.26)
U2 = (12 -f- p cos a) sin a sin <p —
ks cos p,
102 2. Rigid body kinematics - basic knowledge

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),yf (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
uir relative to the axis system Ox'y'z', i.e. the relative angular
velocity. The system of axes Mx*y*z* is fixed in the rigid body S.
The motion of a generic point N characterised by the position
vector p —MI^ 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 uie and is called the
translational angular velocity. It was denoted as UJ 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 v e 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
v e = v 0 + w e 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
2.17 Composition of motions of a rigid body 103

or

wA = v0 + we x r' + vr + (<jje + ur) x p.

(2.17.1)
The first three terms represent the absolute velocity of point M
VA = V + (<jje +
a ur) x p.

(2.17.2)
In particular, let points O, O and M coincide. Then va = 0 and
VA = p = (ue + ur) x p.

(2.17.3)
But in this case, forgetting for the time being about the
vA=uAxp,

(2.17.4)
where LJ denotes the angular velocity with respect to axes
A

Oxyz which implies that CJ is the absolute angular velocity.


A

Comparing eqs. (3) and (4) we conclude that because of


UJ A = uje + ur. (2.17.5)

The result obtained is the theorem of addition of angular


velocities. Moreover, if we set p = i* in eq. (3), i* being unit
vector of the s — th axis of system Mx*y*z* we obtain
di*
= K + wr) xi;^Ax ij,
(2.17.6)
which also follows from eqs. (5) and (7.5).
Vector u:e is naturally prescribed by its projections on axes
Ox'y'z' whilst vector LJr is prescribed by projections on axes
Mx*y*z*. For this reason, the angular velocity vector is
3
eA = UA — (Wesi-s + ^rs*K)
s
=13
= (c5 i(, + ^rs*^) + X LOe + (uJe + U>r) X U>r,
es

s= 1
or
£A — £e + x (2.17.7)
104 2. Rigid body kinematics - basic knowledge

where er denotes the vector whose projections on axes Mx*y*z*


are equal to the time derivatives of the projections of the
relative angular velocity ur on the same axes.
Taking now the derivative of expression (2) with respect to
time we determine a formula for the absolute acceleration of
point N. Differentiating va was performed in Sec. 2.14 and gave
formula (14.13) for absolute acceleration wa of point N. In this
formula one must replace u and e by cje and £e, respectively. The
time-derivative of the second term in eq. (2) remains to be
taken. Using eqs. (7) and (3) we obtain
[(cje + U ) X p] = (ee + £ + LO X LO ) xp+ (cje + LO ) X [(cJe + LO )
r r e r r r

X p] = £ X p + UJ X (uj X p) + (u X U) ) X p + UJ X (u X p)
e e e e r e r

+ UJ X (u X p) + £ X p + u X (cjr X p) .
r e r r

Noticing that
(cje X UJ r ) X p + CJr X (cje x p) = CJ X (ur x p ), e

we can write the previous expression in the form


[(v e X UJ r ) xpf = £e X p + Uj e X (u e X p) + £ r X p+
C + 2cJe X (cjr X p) . O r X (uj r X p)

The final expression for takes the following form


WA = {w0 + se X (r' + p) + cje x [c^e x (r' + p)]}

(2.17.8)
+ {wr + £r x p + ujr x (cor x p)} + {2cje x [vr + (cor x p)]} .
The quantity
wE = w0 + eex (r' + p) + c^e x [ue 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
wu = wr + £r x p + ur x (cur x p)
(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 of the pole M, whereas the second and
the third ones comprise rotational and centripetal
(2.17.11)
Vi? = vr + uj r x p.
2.18 Motion of the natural trihedron of a spatial
105curve

FIGURE 2.17.
Because of this, the third component
WCor = 2u;e x [vr 4- (u>r x p)\ = 2ue x vR

(2.17.12)
represents the Coriolis acceleration of point N. This yields the
theorem of composition of accelerations
W J 4=W£:-fW f i -l- 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 MQ.
The unit vector r that is tangent to the curve is given by the
following equality

dr
T
da' (2.18.1)

Let T\ — r+dr denote the tangent vector at point M' which is


infinitesimally close to point M. When we place vector T\ to
point M as shown in Fig. 2.17 then vectors r and T\ determine
the tangent contact plane II at point M. An infinitesimal vector dr is
perpendicular to r and directed to the concave side of the
curve. Its value \dr\ = e defines an infinitesimal angle between
r and T\ and the ratio
\dr\ _ e _ 1 da
da p
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
di~
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 = r x n, see Fig. 2.17.
Thus, an orthogonal trihedron r, 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 Mi\f of this point
relative to the trihedron origin. dr .as r' and. its projections on the
VQ = r =—a — TO
da
designate the velocity of the trihedron origin. The velocity of
point N is then
. ./ . / _ \® f dr dn db\ .
v = r + r = TO + (air + a2n+a3b) = r + a\— + a2-—ba3 — a.
\ da da da)
(2.18.3)
Alternatively, due to the formula for the velocity of a point in
the rigid body
v = vo + dxr'

(2.18.4)
where 0 denotes the angular velocity vector of the natural
trihedron whilst moving along the curve.
As ai,a2,a3 are arbitrary, a comparison of relationships (3)
and (4) enables the following three equations to be
determined
a^- = f}3n - 02b, a^ — Qib — QsT, a^- = 02r - fiin. (2.18.5) da da
da
It follows from the first equation and (2) that
f^2 = 0, Qs = —.
P
The value of is determined from the third equation in (5)

o •dh • dn • n =ar-
dn\ (2.18.6)
Hi = ■n— TX
da nx
sj-
da
2.18 Motion of the natural trihedron of a spatial curve 107

It is easy to obtain this result by differentiating b = r x n and


accounting for eq. (2) which yields
db dn

The value
z’(a)
1 ( dn\ 2 {dr CPT\ 2 x' (a) y' (V z
" ip)
x"(a) y"W) z"' (a)
T = T
{nXd^)=pT- {toXdZ)=P x'" (a) y ' " ( * ) (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 dr \/\x" (f)] + W (o')] + [z" (a)}2.
2 2

P da (2.18.8)

Taking into account eqs. (5)-(7) we come to the well-known


Frenet formulae dr n dn b r db n ,
.
7 7 = 7 ' 5 ? = r - 7 ' 37 “ " r <2'18!»
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 r of

Mbl =v=
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 Mx*y*z* whose position with respect to
the fixed axes 0^rj( is described by angles A, //, v. The
directions of axes Mx* and r coincide, thus plane My*z* and
the normal plane of the trajectory coincide also. Figure 2.18
shows the rotation through angle x about r that makes the unit
vectors n and b coincident with axes My* and Mz*. Then,
designating the angular velocity vector of the velocity axes as
u? = ft + rx = il
a. .^cr . a
+ + i2 — sin x + i3 — (2.18.11)
X cos x,
108 2. Rigid body kinematics - basic knowledge

FIGURE 2.18.
i* being the unit vectors of the velocity axes. Alternatively, the
definitions of angles A, /i, v yields the following expressions
for the projections u;'* of vector ur on the velocity axes i*
0/1*
= A sin /x + z>, '
^2* A cos/i cos ^ + /isinz/,
=
(2.18.12)
>
* = — Acos/isinz/+/icoszA >
We obtain three equations
a
T + X = A sin fi + z>, (2.18.13)

— sin % = A cos fi cos z/ + /i sin z/, (2.18.14


P )

— cos x = —A cos fi sin v + /i cos v. (2.18.15


P
)
From the latter two equations we find
a [~2

fi,
= V Li +7A
* ~
cos^
P (2.18.16

cos (x - v) = Pv = — -r P
cos2 /i
A cos fi A cos fi (2.18.17)
sin (x-u)= p-
(1 + A cos2 „
2.18 Motion of the natural trihedron of a spatial
109curve

Differentiating the first equation in (17), making use of the


second equation and the expression for x ~ v from (13), we
arrive at the following relationship

(2.18.18)

Formulae (16) and (18) determine the motion of the natural


trihedron for prescribed functions of time A (t) and 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
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 O 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°, that is counterclockwise for a
right-handed coordinate system.
Let us consider the position vector OA? = p before the
rotation. After the rotation it takes position OMf = 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 parameters, i.e. angle \ and axis e, and
vector p.
It is clear that either vector p and pf 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)
A. I. Lurie, Analytical Mechanics © Springer-
Verlag Berlin Heidelberg 2002
112 3. Theory of finite rotations of rigid bodies

FIGURE 3.1.
What remains is to observe a change in the component
perpendicular to the axis. Before the turn it is
Po = P ~ e e P i (3.1.3)

whilst after the turn it is


PQ = p' - ee • p' = f! — ee • p.

(3.1.4)
Figure 3.2 provides a view from the end of vector e showing
that

oj? = 0^ + S M ' or Po = ( P o

Observing that vector SM' has the magnitude 0\S tan — and the
we obtain
11 x
Po= 2 +
+2 ex

p +p
o) tan
2
or
X X
Po -
e
X Po tan - = Po + e x p0 tan -.
Replacing in the latter equation p'0 and p0 respectively by p'
and p due to eqs. (3) and (4), Rodrigues formula is produced
y y
p'-exp' tan ^=p + ex ptan
(3.1.5)
3.1 Rodrigues formula and the vector of finite113
rotation

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
e x (e x p') = ee • p — pi
we obtain
p!
X X
tan ^ + e x p ' = —ptan — + e x p + 2ee • ptan
X (3.1.6)
Z Z Z
Eliminating vector e x p ' from eqs. (5) and (6) we
P’ = have p ^1 — tan2 - j + 2e x ptan ^ + 2ee • ptan2 ^ .
1 +tan2 —
(3.1.7)

This formula can be recast as


follows
p H---------v etan
77 x p + etan ~ fe tan ^ • p) — ptan2 ^ .
1 + tan
1 ^2

(3.1.8)
The two latter terms comprise the following product
e tan ^ x tan ^ x p'j ,
thus formula (7) can be rewritten in the form
2e tan \ / y \
p' = pH---------x (p + etan^ x p) .
-------------------(3.1.9)
l + tan |
2 V 2 }

We introduce into consideration vector 0 whose direction


coincides with
unit vector e of the rotation axis and whose magnitude is
0=\O\ = 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

(3 1 11)
P
' = P +
YT^0X {p+\0xp)' --
The introduced vector
1 = 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 = ii cos a + 12 cos f3 + 13 cos 7,

a, f3 and 7 being angles between e and is , we are entitled to


write
1 = 0iii + $212 + $313?
(3.1.13)
where the quantities
6\ = 2tan ^ cosa, 62 = 2tan ^ cos/?, 6)3 = 2tan 7^ cos7
(3.1.14)
comprise projections of vector 6 on axes Oxyz. However it
would be erroneous to treat notation (13) as a statement that
finite rotation described by vector 0 can be achieved by means

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 6 performs the rotation of axes Oxyz into
Ox'y'z', thus taking p = \s we will have p'= i's and by virtue of (1.11)

= is +
2 /1 (3.2.1)
^ Y+W6x lis + 2dxii

Rotation —0 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)
ex - -0 x i'. (3.2.2)
1 + \02
3.2 Parameters of Rodrigues and Hamilton 115

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


e\'s = 0is, (3.2.3)

i. e. projections of the rotation vector on the


corresponding axes of coordinate systems Oxyz and Ox'y' z' are
equal to each other
6s=es (8 = 1 , 2, 3 ) . (3.2.4)
Expression (1.13) can be written in either of the following
two forms
0 = 0i\^ + $2*2 4" $313 — 0\\\ + $2*2 4" $313-

(3.2.5) h.
s
= (3.2.6)
2 A0
The proportionality factor is chosen under the following
2
x 0 + xi + x22 + xl = 1. condition
(3.2.7)
Its geometrical meaning
l + l#2 =is1 +
easy
tan2^to= —
understand.
1z—. Due to eq.
4 2 cos x

On the other hand, it follows from eqs. (6) and (7) that
2 + -0 2 = 1 + - (0\ + Q\ + Ol) = 1 + (A? 4- \l +
A) = ~.
-2
4
3 2 ^ An An

Comparing these equations we


have
Xy
1
(3.2.8)
AQ = cos — = \J + 4 02
l

Thus, the rotation vector 6 is determined by the four Rodrigues-


Hamilton's parameters

X X X X
Ai
—, = cos a sin —, A2 = cos (3 sin —, A3 = cos 7 sin —, Ao = cos
(3.2.9
)
subject to a single condition (7). Here <a, /?, 7 are angles of
rotation relative
to the basis axes (coinciding in coordinate systems Oxyz and
93 =
93
o = T- Y] Asis = ° 7^i
A vA X]Asi«’ (3.2.10)
° T^i

3/ 3\
is — is T 2 ^ ^ Xfif x 1 Aois T ^ ^ ^gig * is I (3.2.11)
•>
t=1 \X9=1 /
v 1l * l
xf Ao + Af - Ag — (A0A3 + A1A2) | 2 (—A0A2 +
2
yf A 3
2 (—A0A3 + Aj^-A^-A? | A 1A3) |
2 ( A 0A 1 + A 2A 3)
zi A2(A
2A1Q
) A2 + | 2 (—AQAI -T A
o + A3 - \ \ - \ i
A3AI) A 3A 2) ||

is = i's - 2 ^ A(ij x A0i's - £ Xqig Xi'8 .


(3.2.12)
t= 1 \ <7=1 /
Recalling the definition of the Levi-Civita symbols and using
the following equality
it X (ig X is) — 1 q&ts isAgj
we can transform relationships (11) and (12) to the form
is = is + 2 | Ao ttsq^tiq + A? Xqiq — is X2q J ,
/ 3 3 3 \

t=1 9=1 9=1

3
3

is — is 2 | Ao ^ ^ Aig A ^ ^ Ah? is ^ V A (3.2.14)


t=1 J 9=1 9=1

These yield expressions for the direction cosines of angles


between the axes of these coordinate systems

&sk — ^ * ifc = $sk ~ 2 A^ + 2Ao ttskXt + 2AsAfc, (3.2.15)


V 9=1 / t=l

<^fcs — is • ifc = Afc (1 — 2 Ag j — 2Ao ttskXt +


2AsAfc. (3.2.16) 9 = 1 V / t=i

Recalling the values of the Levi-Civita symbols and making


use of relationship (7) we obtain the table of direction cosines.
It is constructed row-wise by means of eq. (15) and column-
wise by eq. (16).
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 coordinate transformation 3 3 3

xs — (A0 AJL A2 A3) xs + 2Ao EE £tsk^t%k T 2AS ^ ^ A/-X/-,


k=1t=1 k=1
(3.2.1
7)

33 3
xs = (A2- A ? - A 2-Al)a:'s-2Ao E E Ctsk^tz'k + 2 As Akx'k-
k=1 t=l k=1
(3.2.1
8)
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’ - 1 ^ 1 q2 x (P’ ~\6x P') (3.2.19)

on axes Ox'y'z'. Equation (19) is an inverse form of the


Rodrigues formula and expresses that rotation —6 makes
vector p' coincident with p, as well as axes Ox'y'z' coincident

3.3 Composition of finite rotations


Let a rigid body, having an immovable point O, be subjected
to rotation 01, i.e. it rotates through an angle \i about the axis
prescribed by the unit vector ei. Afterwards the body is
subjected to a second rotation O2 characterised by an angle x2
an
d the unit vector e2- The vectors ei and e 2 are assumed not
to be related to the body. The angle (less than 180°) between
these vectors is denoted by then
ei*e2 = cos$, |ei x e2| = sin$.
(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 0, angle x and axis e, is required to be
expressed in terms of vectors 6\ and 02-
We construct a spherical triangle ABC on the unit sphere
about O as shown in Fig. 3.3a. Its vertices coincide with the
118 3. Theory of finite rotations of rigid bodies

€'

BC with angles |xi and \x2 relative to arc AB as shown in Fig.


3.3.
Indeed, rotation 0\ carries point C to Cr whilst rotation 62 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
02 followed by rotation 0i does not result in point C but in C'.
The direction of the axis of the resultant rotation is now
defined. What
is left is
determineangle x- To do thiswe now turn to Fig. 3.3b.Point
A remaining fixedatthe firstrotation takes position
after the second (3.3.
2)
rotation. This is also the position of this point at the resultant
rotation 0
through angle x- Thus the angle at vertex C of the spherical (3.3.
triangle ABC 3)
is equal to ^(27T — x) — ^ — ^x-
Now we derive the formulae corresponding to the above
geometric con- (3.3.
struction. We will use formulae of spherical trigonometry
cos c = cos a cos b + sin a sin b cos C,

sin a sin b sin c


sin A sin B sin C ’
3.3 Composition of finite rotations 119

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

cos | = COS ^ cos y (^1 - ■ 02^) ■


(3.3.6)

We derived expression for the angle of the resultant

e = aei + (3e2 + ye2 x ei,


(3.3.7)
where the coefficients are to be determined. Coefficient 7 is
positive as vector e (see Fig. 3.3) deviates from the plane of
vectors ei and e2 to the side of e2 x e i. The case 7 < 0
corresponds to the case in which rotation 02 precedes rotation
0\.
Since e is a unit vector, eq. (7) yields
or
\/l — a2 - (32 — 2af3 cos $. (3.3.8)
7 = sin<I> n (D V

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


first calculate the scalar product of eq. (7) and e 2 x (e2 x e i ) ,
and second the scalar product of eq. (7) and ei x (e 2 x e i ).
Performing these operations we obtain
e • [e2 x (e2 x ei)] = aei • [e2 x (e2 x e i ) ],

e • [ei x (e2 x ei)] = 0e2 • [ei x (e2 x e x ) ] .


Then by means of Fig. 3.3a and formulae (1) and (2) we
obtain
e • [e2 x (e2 x e i ) ] = e • e2ei • e2 - e • ei
= cos a cos# — cos 6 = — sin a sin
e- [ei x (e2 x e x ) ] = e • e2 — e • eiei • e2
= cos a — cos b cos = sin b sin $ cos ,

ei • [e2 x (e2 x ei)] = (ei x e2) • (e2 x e i ) = - sin2

G2 • [ei x (e2 x ei)] = sin2 <E>.


120 3. Theory of finite rotations of rigid bodies

Now making use of eq. (3)


we find
• X2
sin a Vo sin Xi
2X 2 4r. P- sin b Xi
-—- cos —
sin
T Xi
a = -—— cos —- = sin$ 2 X COS 2
sin 2 X cos 2
sin 8111
—2 2
(3.3.9)
By means of eq. (3) we
obtain
7 = ( • 2 X • 2 Xl 2 X2
1 / . 7 sin — — sin — cos — —
---------T
•2 X2 1/2
sin' X2 2X1 o . Xi • X2 Xi X2 \
2
— cos —----2 sin — sm — cos —f
27
.2 X
■ /V, X
Replacing here sin^ ^ = 1 — cos2 — by its expression due to eq.
(5) we find
sin.—Xism
• X-f2
22
•X (3.3.10)
sm-
Inserting expressions (9) and (10) into eq. (7) we have

—y (ei sin cos ^ + e2 sin y cos y + e2 x ex sin y sin y ^ ,

(3.3.1
which by means of the definition for the rotation vector can
1) be
cast in the form
cos Xi
— cos
2 2
X—
2
0— 0\ O2 -02 X 01 (3.3.12)
cosX
By virtue of eq. (6) we arrive at the formula expressing the rule
of composition of finite rotations

0— ^ 10i 02 + ^2 + -02 x 0i^ •


(3.3.13)
The presence of the term 02 x 0i indicates that finite rotations
are not commutative. That is, the resultant rotation of 02
followed by 0i is given by the following formula

0' = 01 T 02 H" 7T01 X 02 (3.3.14)


1 - \Bi • 02
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
0i = 2etan 02 = 2etan

we have 0
2e + tan = 2e tan X i +X 2
X 2 1( f
tan
= 0' ,X i ,
— tan — tan — 22
(3.3.15)
It is easy to obtain expressions for the Rodrigues-Hamilton
parameters of the resultant rotation (denoted by ZA), z'l, ^ 2, ^3)
m
terms of those parameters Ao, Ai, A2, A3 for the first rotation
and /i0, /ilJ fi2, M3 for the second rotation. First by means of eqs.
(2.8) and (13) we calculate the following value
1
1 + V = l + —-------i--------- - (oj 0l 2d -e + ^ |02 X 0j|z ) .
+ + 1 2

4(1--02)'

Since
102 X 0j|2 = Q\Q\ - (0! • 02)Z

we obtain
1 (1 + l A ) (1 + & )
" 2 (1 - \0i • 2’
'0 02) AOMO - E
As/i.
and thus we can adopt that
3

= A0/i0 -EX1A MS- ^0 S


(3.3.16)
S=
1
Now by means of eq. (13) we obtain

Vs_ A0M0 (rA + ^ + y^-EE


vo \ o Mo AQMO ~it^i
A0M0 ^2 As/is
s= 1

or
3 3

V — A sMo + Ms Ao + EE (3.3.17)
r= 1 t= 1
122 3. Theory of finite rotations of rigid bodies

From eqs. (16) and (17) we obtain the following system of formulae
z'o — —
Ai/ij —^2/^2— A /i3, 3

vi = Ai/JL0 +Ao/ii+A /i - (3.3.18)


3 2

A /i3, ^
2

^2 = A /i + A /i + Ai/i - A /i1?
2 0 0 2 3 3

It is worthwhile noting that these expressions present


formulae for multiplication of quaternions.

3.4 Subtraction of finite rotations


Finite rotations possess the property of associativity. This
means that the sequence of rotations 0 i , 0 2 , 0 3 can be
performed by composition of the resultant rotation of 0i and 0 2
followed by 63 or by composition of 0\ followed by the resultant
rotation of 62 and 63.
With this in mind, let us consider the following sequence of
rotations: —0i,0i, 02 which is obviously equivalent to 0 2. This
rotation is equivalent to the sequence of rotations —0\ and 6.
Therefore, replacing in eq. (3.13) vectors 0 , 0 1 ? 0 2 by 02, — 01,0,
respectively, we arrive at the formula for subtraction of
rotations

(341)
e
* = TTWii(e-e' +
\tl'xe)’
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 — 0 2 and —0\ is
equivalent to the resultant vector —0. For this reason, using
eq. (1) we obtain

-"■ = TT^(-<' + 92 + ^xS)


or

(3 42)
e
^TTWFXe-e2 + \e^^ '

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 assumption we arrive at conclusions
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
6>i = 2eitan^, 02=2e2tan^-

(3.5.1)
z z
is equivalent to the sequence
0\ = 02 = 2e2 tan d'2 = 2e[ tan
(3.5.2)
where stands for unit vector into which rotation 62 carries e.
This vector is determined by eq. (1.11)
ei=ei + -—T7202 X (ei + \o2 X ei ) ,
(3.5.3)
1 + 4^2 \ Z
/
and then

62=61 + ----------
X
o^2 X
, 1 - \6 X • 62
*2=<h + 02X0. (3.5.4)
1 + 4^2

Now we will construct the resultant rotation 6 f of the rotation


sequence
(2) . Equation (4) leads to
0'i • 0' = 62 • 0'2 = 02 • 0i. (3.5.5)
Then by means of eq. (3.113) we obtain

0' =

1 02 + 01 H-1 * 12 ^2 X 0+ 2

+ 4^2
1 - {01 • 02
>< ^2 + 1 1 x (*2 X 0) X 02

Then we add the vector


—0i x 02 + —02 x 0
lj
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

0' =0+ 1 62 X
(1 - \ex • o2) (1 + &)

Replacing 6 in the square brackets by the corresponding


expression from eq. (3.13) we obtain the vector

which is collinear to O2 . For this reason, substitution into eq. (6) yields
0 = 0 ', (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

(3.6.1)
through angles $, p about axis Oz, the nodal line n and axis Oz',
respectively. 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 62 brings vector i3
into coincidence with vector i's. Thus, by virtue of the above
theorem, the sequence of rotations (1) is equivalent to the
following sequence
(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 + p, i.e.
(3.6.3)
2
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),
s 3

n o( , $ ./ ^ + P, $ p + lb
6 = 2 I n tan - + 13 tan + n tan - tan -—- (3.6.4)
\222
Projecting this equation on axes i'l5 i2, i's we have

9 = 2 ij tan — ^cos p + sinp tan — ^ ^ + (3.6.5)


, # / . p + ?/A xp +
<z?
i2 tan - I - sin + cos ip tan —-— ) + 13 tan
'd 'ip — p . %p — p . p + %p
ii tan — cos ——+ i2 tan - sin ■ +13 sin ■
p + ip
cos

In this expression the coefficients of i' are projections of the


rotation vector 6 on axes Ox'y'z' as well as on axes Oxyz. The
magnitude of the rotation vector x is determined by eq. (2.8).
This leads to
1 + -7 O = (1 + tan2 x ) (1 + tan2
2

and

cos2 -X 2=
d2
= cos^
cos - -
V +^
cos^ ——, Xcos $
| p + 'Ip
cos

The sign of parameter Ao can be taken arbitrarily. For this


reason we obtain from eqs. (5), (2.6) and (2.8) that
. d p + ip . . $ 'ip — p \
Ao =cos — cos —-—, Ai =
sm — cos —-—, (3.6.6)
Z Z z z
. $ . Ip — p - d.p + tp
A2 = sm - sm —-—, A3 =
cos - sm —-—.
These are expressions for the Rodrigues-Hamilton’s
parameters in terms of Euler’s angles.

1.7 Applications of formula for finite rotation


1.7.1 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 e i , e 2, e 3 denote
126 3. Theory of finite rotations of rigid bodies

FIGURE 3.4.
the trihedron of axes bound to the platform of the suspension,
ei being directed along the rotation axis of the outer gimbal.
Also, ii, i2, h (h = ei) and i'l5 i2, i3 (12 = i^) 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 i3.
Initially all trihedrons coincide.
It is necessary to prove the equivalence of the sequences
(3.7.1)

and
(3.7.2)

since i[ is the position of vector ii after rotation


0\ — #2- The resultant rotation of sequence (1)
is given by
(3.7.3)

whereas that of sequence (2) is

(3.7.4)

It is necessary to prove whether the equality


.
ii + i2 x ii tan — = ii + i3 tan —

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
ii = ii cos (3 + i3 sin /?, i'2 x ii = ii sin /? - i'3 cos /?,
3.7 Applications of formula for finite rotation
127

u
h

and the problem is reduced to proving the following

identities cos (3 + sin (3 tan ^ = 1,— tan ^

cos (3 + sin (3 = tan


P
a parameters a P
Rodrigues-Hamilton’s
v0 = cos - cos -, V\ = sinof- rotation
cos -, 6 equal
(3.7.5)
a . (3 . a . (3
v2 = cos — sin-, i/3 = sin — s i n - .

The same expression can be obtained from eq. (3.18) when


we notice that, due to eq. (2),
\ & \n\ •P
A0 = cos —, Ai = 0, A2 = s i n - , A3 = 0,
a •an /x3 = 0.
/x0 = COS-, Mi = Sin-, /i2 — 0,
is easy to
By means of Table 1 of this chapter and reconstruct
eq. (5) it
matrix (2.6.2).
1.7.2 Rotation of the geocentric system of axes
The definition of this system of axes is given in Sec. 2.15. Let
$0 and Ao be the northern latitude and eastern longitude of
the origin of the system at an initial time instant t = to where $
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


— <F AA
1 = 2e2 tan —-—, 0 02 = 2etan---. (3.7.6)
Z Z
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 AA = 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 AQ
from the point with latitude <3>o to the point with latitude $
and the second rotation along the parallel of latitude <F. By
virtue of the theorem on commutative rotations the sequence
(3.7.7)

where e2 denotes the unit vector perpendicular to the


meridian plane A pointing west. Rotation 0 2 = 0[ carries e2
into e2. Sequence (7) first implies displacement along the
parallel of latitude <3>o and then along the meridian A to the
parallel <F. Since the final position of the trihedron does not
depend on the order of our mental constructions, this
example
We havereadily illustrates the theorem on commutative

(3.7.8)

The identity 0 = 0' can be proved by analogy with the previous


example. Since

e = ei cos + e3sin$0, e x e2 = -ei sin$0 + e3<F0,


where e3 denotes the unit vector of the upward vertical and ei
is directed north, we obtain

0
2
(3.7.9)
2 2
3.7 Applications of formula for finite rotation 129

The Rodrigues-Hamilton parameters are


as follows
$ - <J>0 AA
A0 = cos —-— cos —,
. <3> + $0 . AA
AI = cos- -- --sin —,
(3.7.10)
, . AA '
A2 = sin--------cos ,
2 2’
. ^ T $o . AA
A3 = sin---------sin .
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 0 at the start point (<3>o, Ao). This
direction is described by an initial (at time instant t = to)
azimuthal angle ao (angle counted between ei and e 2) and an
initial ascent angle /?Q 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'z' at
the actual place of latitude <3> and longitude A at time instant
t.
The coordinates of the end of unit vector of the rotor axis
are given by
in coordinate systems Oxyz and Ox'y'z', respectively. Using eq.
(2.17) we obtain
/ 9 9 9 9 v
cos a cos /3 = (A0 + Ax — A2 - A3) cos ao cos /30+
2(A0A3 + AIA2) sina0 cos/30 + 2 (-A0A2 + AiA3)sin/30,
sin a cos (3 = 2 (—AoA3 + AIA2) cosao cos (30+ (3.7.11)
(A Q + A — A^ — A§) sin ao cos f3§ + 2 (AoAi + A A3)
2 2

sin / 3g,
sin/? = 2 (A A + A I A 3 ) cosa0 cos/?0+
0 2

where Ao, Ai, A2, A3 are due to eq. (10). These formulae
simplify in particular cases. For example, provided that the
rotor axis was initially on the horizontal plane and pointing
east we have ao = — ^7r,/? = 0 and formulae
(11) take the form
cos a cos (3 = — sin AA sin <3>, sin a cos (3.7.12)
(3 = — cos AA, sin /3 = sin A A cos <3>.
130 3. Theory of finite rotations of rigid bodies

When the rotor axis was initially on the plane of the meridian
(CKQ = 0) then calculation with the help of eq. ( 11) yields

cos a cos (3 = cos cos (4>o — Po) + cos AA sin $ sin ($o
~ Po) > 1 sin a: cos/? = — sin AAsin (4>o — Po)>

>
sin P — sin cos (4>o — Po) — cos AA cos sin (4>o — Po) • J
(3.7.1
3)
The most simple case is that is which the rotor axis is
directed to the north, then 4>o = Po and, by virtue of eq. (13),
we obtain
cos a cos P = cos 4>, sin a cos P — 0, sin P = sin 4>,
which implies a = 0, P = 4>, i.e. the rotor axis maintains the

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 0
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 6'0 from the actual position or by subjecting the body
to a rotation 0 + 60 from the initial position. Here the
infinitesimal vector 60 is an increment in 0 due to the addition
of an infinitesimal rotation such that 0 + 60 is the resultant
rotation of two consequent rotations: finite 0 and infinitesimal
6f0. 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
6'e =-----V-5- {60+-9 x 60) .
---------(3.8.1)
i + \e2 V 2 ) v
'
6'0 + \#0 x 0 + -QO-6'O
60 2 4
6'e(i + \?)-lex(s' 0 + 16'0 x 0 (3.8.2)

The angular velocity vector u is related to the vector of


infinitesimal rotation by equality udt = 6'0. Making use of eq. (1)
yields the equation
3.8 Expressions for the angular velocity vector in terms of finite rotation 131
relating u to the vector of finite rotation

C0 — :—( 0 H“ —0 X 0
------- (3.8.3)
l + \ 2

The time derivative of the rotation vector is expressed in terms


of the angular velocity vector by means of the formula
•11
0 — u?T—oj x 0-\-—00 • u). (3.8.4)
2 4

Differentiating expression (2.10) for 0 and taking into


account eq. (2.8) we obtain
0
= 2 ^AoAi — AiAo^ ii + ^AOA2 — A2Ao^ 12 + ^AoA3 —
AsAo^ i3
l + ±02
(3.8.5)
where ii,i2,i3 are the unit vectors of the fixed axes Oxyz. Due to
eq. (3)
the projections of vector uo on these axes are
£4 = 2 ^AoAi — A1A0 + A2A3 — ASA2^ ,
1) 2 = 2 ^AOA2 —
A2A0 + A3A1 — AIA3^ , > (3.8.6)
Cb3 — 2 ^AoA3 — A3A0 + A1A2 — A2/\I^ .
Projecting now eq. (2) on axes Oxyz yields the relationships
2 ^AoAi — AiAo'j = Cb\ (AQ + A2) + 0)2 (A0A3 + A1A2) + Cb% (—A0A2 +
A1A3)
2 f A0A2 — A2A0J = Cb 1 (—A0A3 + A1A2) + (AQ + A2) + Cb% (A0A1 +
A2A3)
2 ^AoA3 — A3A0J = Cb\ (A0A2 + A1A3) + Cb2 (—A0A1 + A2A3) +.0)3
(AQ + A3)

(3.8.7
)
Using constraint equation (2.7) and its consequence in the
(3.8.9)
form
A0Ao + A1A1 + A2A2 + A3A3 = 0, (3.8.8)
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
uj\ — 2 ^AoAi — A1A0 + A2A3 —
Co 2 — 2 A3A2^ , (3.8.10)
C03 = 2 ^AOA2 — A2A0 + A3A1 — AIA3^ ,
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 —0. This means that the signs of Ai, A2, A3 and co
changes whilst the
sign of Ao is unchanged. Expressions for the time-derivatives
of Rodrigues- (3.8.11)
Hamilton’s parameters in terms of cos take
the form
2Ao = — (CLJIAI + co2A2 + CO3X3),
2Ai = co 1A0 + ^3 A2 — co 2 A3,
2A2 = CO 2A0 + CO 1A3 — CO 3A1,
A0 + CO2X1 — CO 1XAA
1„CJ2 23$-$0
2A13, = 2-
1 H—-0 = cos —-— cos
2‘
Differentiating with respect to time and taking into account
that vectors e and e2 do not move we obtain
$ AA
e e2 + AA e +
cos^ cos^
AA $ - $0 $ AA \
■ tan- -----b ——r=— tan —— e*,
AA $ - $0 2
cosz cos^ J
where e* = e x e2 is a unit vector directed outward along the
radius of the parallel circle, e 2,e*,e forming a right-handed
trihedron of unit vectors. Inserting this expression into eq. (3)
we have
co = eAA + e 2<i> cos AA + e*<I> sin AA.

(3.8.12)
Projecting this equality on the axes of the geocentric system
we obtain
3.9 Cayley-Klein’s parameters
133

where ei,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 =Ue designates the angular velocity of the earth.
Expressions (12) and (13) can be also represented in the
form
a; = eAA + e'2<& = U + A (e^ cos $ + e3 sin $) + e^,
where are the trihedron of the geocentric system of
the actual

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 EQ of
unit radius. Points in the plane are projected
stereographically from the pole N on the sphere EQ, that is
points iV, M and M' lie along a straight line, see Fig. 3.6. The
fixed axes are
taken so that the origin lies at the
centre
of the sphere, axes and £ 2 coincide with axes Ox and Oy,
134 3. Theory of finite rotations of rigid bodies

line NMMf is expressed as follows


y_ = 1
z = x + iy = £i + i£ 2
=
(3.9.1)
fi 1-^3 ’
the quantities £1? £2 and £3 being related by the equation for the unit sphere
£?+ £! + £§ = i- (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
1-£| B^ + Ch
A (x2 + y2) - (Bx + Cy)
(1-&)2 !-«3

is transformed into a circle equation


2A J -Y B2 + C 2
+ \y 2A 1 +
4 A2
in the plane
-C{2 = 0
passing through the origin O of the coordinate system.
B2 PC2
This circle contains the sphere centre since the radius1+ R
4 A2
of the circle is greater than the distance between the centres
B2 + C2
of the circle . It is easy to show (however
4 A2
we omit
and the the proof) which
sphere that stereographic
is 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 N
and S. The unit circle \z\ = 1 corresponds to the equator, points
within the unit circle (i.e. \z\ < 1) correspond to points on the
lower hemisphere (£3 < 0) and points outside the unit circle (W
> i) correspond to points on the upper hemisphere (£3 > 0).
The origin z = 0 and the infinite point of plane P correspond to
the poles S and N, respectively.
We turn now to the case of a rigid body with a fixed point O
and imagine a sphere E circumscribed about O and rigidly
bounded to the body and axes Ox 1X2X3. During the body
rotation the sphere E slides over the fixed sphere Eo- The
rotation changes the great circle T of the sphere E into circle
3.9 Cayley-Klein’s parameters 135

great circles T and T'. This change of 7 to 7' corresponds to a rotation of the body
which is equivalent to a rotation of the sphere E. Under the above change the
complex-valued coordinates of the points of circles 7 and 7' are related to each other by
means of a rational transformation
. az T Q
z =-------- (3.9.3)
7 z+o

The coefficients a,/3,j,6 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

aS- 07 = 1. (3.9.4)

The quantities a, (3,7,6 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 zx,z2, z3 of
circle 7 to three points z[,z'2, z'3 of circle 7' can be written in the form

Z Z2 Z3 Zl
_ Z 2 3 Z'l
Z Z
/g g
z - Zi 23 - z2 z' - z[ z'3 - z'z V••j

The rotation vector 0 intersects sphere E0 at point P, whose coordinates are


proportional to the Rodrigues parameters Ai, A2, A3 and at a diametrically 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

A,
£1 =± £2 =±
A3

\A- ^0 1-A
2
0
In view of eq. (1) the corresponding points P' and Q' in plane P have the coordinates

Ai + 1X2 Ai + i\2
=----------, ^3 =- - -•;= ------.
1 — A — A3
0 y 1 — AQ + A3
Since points P and Q does not move under the body rotation, we can set

£2 = z'2, z3 = 4
in the coordinate transformation (5). Point z\ = oc corresponded to the
pole N before the rotation. The coordinates 442^3 of this point after
rotation are

£1 — 2 (A A + A3A1), ^2
0 2 2 (—AQAI + ASA2) , £3 — AQ + A3 — A^ — A2,
136 3. Theory of finite rotations of rigid bodies

which follows from eq. (2.18) by taking x\ = x2 = 0. Taking into


account eq. (2.7) we obtain, by means of eq. ( 1), that
A3 —
zAp AI
Substituting — 1X2

Z-i = z'2 Ai + ZA2 Ai + ZA2


^/l-Ag-Aa’ \A ~ AQ+A3

A3 — zAp AI
Z\ — 00,
— Z A2
into eq. (5) we have
z (A3 — zAo) + Ai + ZA2 z (Ai — ZA2)
— (A3 + ZAQ) ’

and thus
a - C(A - zA0), (3 = C(Ai + zA2), 7 - C(Ai - zA2), S =-C (A3 + zA0).
3

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
OL — Ao + 2A3, (3 — —A2 + zAi, 7 — A2 + zAi, 6 = Ao — ZA3.
(3.9.6)
One can see that a and S, as well as (3 and 7 are complex
conjugates. Moreover, \a\2 + \/3\2 = |7|2 + |<S|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 = cos - exp It -y- J , (3
1 sin — exp 1
22 > (3.9.7)
■■
7 = z sin — exp

S = cos — exp l-i—-—


The transformation of coordinates (xi,X 2,x3) into (£1, ^>^3)?
expressed
by eq. (3), can be written in the form
=
( ^ s') (Xl’X2>x3)-
(3.9.8)
z— (3.9.9)
However it follows from eq. (3) that
-6z‘ + (3
3.9 Cayley-Klein’s parameters 137

Thus, the inverse transformation, i.e. transformation of (^ 1,^2^3)


(xi,X2,xs) is as follows

(x1,x2,x3) = P_a ^ (£I,£2,£3)- (3.9.10)

An explicit expression for transformation (8) due to eq. (1) is


£1 + i£2 _ a(xi +1x2) + (3(1 - x3) (3.9.11)
l-£3 7 (xi + ix2) + <5(1 - x3)
The complex conjugate is then
£1 - i£2 _ s(xi ~ix2) ~ 7 (1 -£3)
(3.9.12)
1 - £3 ~(3(xi - ix2) + a(l - x3)

Multiplying eqs. (11) and (12) we obtain by virtue of eqs. (2) and (4) that
1 + £3_ 1 + [(a<5 + (3j) x3 - (cry - (36) x\ - (07 + (36) ix2] 1 — £3 1
- [(a<5 + /?7) x3 - (07 — (36) x\ — (a7 + (36) ix2\
and conclude that the expression in the square brackets is £ 3.
Using this
expression for £3 we obtain + i£2 and £j — *£2 an(l arrive at the
trans-
formation formulae
(3.9.13)
£1 + *£2
=
(^1 + ^2) - (3 (x\ — ix2) — 2a(3x3,
a2 2

£1 - *£2 = _72 (^1 + *^2) + <52 (xi - 1x2) + 2^6x3,


The inverse transformation is obtained by using eq. (10)
%i + ix2 — 62 (£j + Z£2) — (32 (£1 *£2) + 26(3£3, 1
xi - ix2 = -7 (£1 + *£2) + “ (£1 ~ *£2) “ 27 £3> /
a
2 2
(3.9.14)
£3 = 7*5 (£1 + *£2)
_
/? (£1 *£2) + ( $ + Pi) £ 3- J
a _ a

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 (3.9.15)
transformations, namely
the transformation
OL\Z + Pi ,
z =---------—
7l-2 + 0l
138 3. Theory of finite rotations of rigid bodies

corresponds to a second rotation


The transformation z into z" is given by

(3.9.16)
a1z + 01 + ^ 7 nz + 612
2
7xz + 61

and describes the resultant rotation 6.


Thus, Cayley-Klein’s parameters of the resultant rotation are
expressed in terms of rotations 0\ and O2 as follows
<*12 = < * 1^2 + 7 I / ? 2 > P12 =
(3.9.17)
+ <5l/^2>

712 = 72^1 + 7i^2, s12 = + M2-

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
a, /?, 7, 6; and a + adt, (3 + /3dt, 7 + 7 dt, 6 + 6dt,

respectively. In eq. (9.17) they can be understood as


parameters aq, f31, ,8\
of the first rotation and parameters a12?/?I2?7I2> £12 of the
resultant rotation. In the case under consideration the second
rotation with parameters a2 , /?2 ^ 725 ^2 is an infinitesimal
rotation 8'Q =tudt whose projections on axes 0£I£2£3 are equal to

Ao Ao Ao
Then with accuracy up the to second order we can adopt

A0 = 1, Ai = A2 = ^(J2 dt, A3 = ^U3dt


(3.10.1)

(3.10.2)
3.11 Determination of a rigid body position from angular velocity 139

Using eq. (3) we obtain

a + adt = ( 1 + ^cl>3dt'j a + 1 (-u>2 + id) 1)7dt, (d


(3.10.3)
+ fddt = | 1 + ~d)3dt^ fd + — (—0)2 + iwj) 8dt

etc. Then we find


XX 'XX
a = -u3a + - (<Di + iu)2) 1, P = -jd)3(d +-(u>i+ id)2) 8,
XX 'XX
7 = “2^37 + 2 (^! ” ^2) a, 5 = ~ + - (a)i - ia)2) 7*
(3.10.
4)
Replacing here the projections us of the angular velocity
vector on axes 0£I£2£3 by their expressions in terms of
projections UJ on axes fixed in the body with the help of
S

formulae (9.14) we obtain


=l
^2~a + \(ui ~iuj2">i3 XUJ3 D 1 < ■ \
—2"P + - (ui + xuj )a,
► (3.10.5) 2

. ^3 1/ . x c -c XUJz c X , . .

7 = -g-7 + - (u)\ - xu2)b, 6 = ——6 + - (cJi + XLU2) 7-

It is easy to construct the inverse relationships


u 1 -f iu)2 = 2i ^<5/3 — a(3^ ,
(3.10.6)
u)i — iuj2 = 2i ^67 — £7^ ,

0)3 = 2i ^7 — = 2x — 7/?^
and

CJI + XLU2 = 2i (^36 — fiS'j ,

ui — iuo2 = 2i (7a — 7a), (3.10.7)


CJ3 = 2i — j3^j — 2x (/3j — aS).

3.11 Determination of a rigid body position from angular


velocity

It is assumed that projections ttq, tt> 25^3 of the angular


velocity vector
of the body rotating about a fixed point on axes fixed in the
body are
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 aq, LL>2, ^3-
A first form of these differential equations is given by
expressions (2.10.1) for derivatives of Euler’s angles in terms
of projections LU of the angular velocity vector
S
$ = (J0\ COS <p — CJ2 sin ^9, '

^ = a (<+7 sin (p + U)2 COS <p), (3.11.1)


>
smv
Certain of the particular cases admit integration of this
system. However 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 7i, 72,73 on axes fixed in the body, i.e. the
direction cosines of angles between
1
e and axes Ox'y'z', we have
S=

As the velocity of the end of vector e is equal to zero, we can


write, due to eq. (2.13.3),

3
e = E ('Vs + x i'J = 0.
S— 1

It follows that the coefficients of vanish which yields a


system of three differential equations 7

7i + ^273 - ^372 = 72 + ^37i - ^173 = 73 + ^172 - ^27i =


0.
(3.11.
2)
7i7i + 7272 + 737S = 0.
and its integral
7I+72+73 = 1- (3-11.3)
Also the initial values 71,72,73 should satisfy this condition.
Thus the problem is reduced to integration of the system of
three differential equations (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 = ii which implies
that vector e is directed along axis Ox. Then we can determine
the integration constants by means of the initial conditions 7° =
aJi, 7° = 0^,73 = a3i satisfying condition (3) where denote
prescribed direction cosines of the angles between axes Ox'y'z'
and axis Ox at the initial time instant. We obtain an,a 2i,a3i
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 parameters. In this case we need to integrate a
system of four differential equations (8.11) having the first
integral (2.7). If we introduce complexvalued 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
first 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
. i&3 1, . . . iujs 1 , 2 luJx x
3 n 4
x
^ = — * + 2 (^2 + i u i ) P , P = — Y P ~ 2 ^ ~ ) (* - )
*i( 0) = l P i ( 0) = 0 (3.11.5)
*2(0) = 0 p2( 0) = 1

are known and satisfy the initial conditions. The solution of


system (10.5) subject to the initial conditions
t= 0 a= a0, = 7o,
0 = I3O, 7 ^ = S0,
(3.11.6)
which satisfy the following relationship
a060 - /?07o = 1
can be cast in the form
a = a0xi + 00^2, 0 = aof>i + 0op2, 7 = 70xri + S0x2, 6 = 7Qpx + S0p2.
(3.11.
8)
All we need is to prove that this solution satisfies condition
(9.4). Recall
that the system of the differential equations
i = P11 (t) x + P12 (t) y, y= P21 (t) x + p22 (t) y
has the Wronskian
D (t) = xiy2- x2yi = D (0) exp t

J {Pn ~\~
142 3. Theory of finite rotations of rigid bodies

But for system (4) pn + P22 = 0 and for the assumed solutions D
(0) = 1. Thus
X1P2 - *2pi = 1,
and, in view of (8),
aS- j3j = (a060 - /?07o) (*iP2 ~ *2Pi) = 1,
which completes the proof.
Actually, it is sufficient to find only one system of particular
solutions , px. The second system is obtained from the
equalities
= -Pi> P2 = ^h ^2
where a bar denotes the complex conjugate. Indeed, K\ and px
are solutions of the system of differential equations obtained
from (4) by replacing % with —i in the right-hand side. In
addition to this, = 1 and px = 0. These are the functions K<I 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 problem 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 EQ, the problem being discussed in
Sec. 3.9. Let £i ,£35 £3 designate the coordinates of a point on
the sphere. Its coordinates relative to axes Ox 1X2X3 are given
by transformations (9.10) and (9.9). Differentiating the latter
equation with1 respect to ftime and taking into account that i' = 0
Sz + (7z' a) - (Sz* + (3) (7z* - a) .
(7 z' — a)
2

Applying formulae (10.6) we have

i =---------------------------- * \u>i + iu + 2Z'CJZ - z


2
/2
(dq - id)2) 1 • (3.12.1)
2i (7z' — a)
Since, due to eq. (9.3),
1 - (72? + P), -—p—-2 = {OLZ + P) (pfz + 6),
(7 z' — a)
7 zf — a
(az + P )2 ,
(7 z' — a)2
3.12 Darboux’s equation 143

we use formulae (9.14) for transformation of


vector uo LO\ 4- icj2 — <5 (ct)i 4- iCo2) — /? (CJ\ — 10J2)
2 2

c^i — = —72 (o)i 4- iCo2) 4- a2 (u>i — 20)2) — ► (3.12.2)


4~ 2/4^0)3,
270:0)3,
(^3 — 7 <5 (07 4“ 1002) ~ pot. (Co\ — 1C02) 4“ (OL6
4“ /?7) ^3?
(3.12.3)
> 2 ' 2
The form of this equation can be simplified. Using the
substitution

z = (exp i JU 03dt^ (3.12.4)

t
*/
C = - (002 ~ iooi) exp [ i U03dt 4- - (002 4- iu\) (2 exp —i oo3dt/■.

(3.12.5)
Next assume
(002 4- iooi) exp ^3dt J = Qq,

so that
Z
_1
12 = q = exp i arg
(OJ2 + —UO^dt
J
iuj{) q = ~-
q
(3.12.6)
-J 0
Now introducing a new independent
variable
z t
1 (3.12.7)
= \J ndt, T:

2
which increases monotonically as t increases, we cast the
differential equa-
tion (5) in the form
(3.12.8)
C = -+q(2,
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
T) = -—i = (3.12.9)

satisfies a linear differential equation of the first order. This


equation is easy to construct. Indeed,

t C -<£ 9 ( C - a

V C- £

However from the differential equations which is the complex


conjugate of eq. (8) we have
c' =q+ i 1 i = l_l
* ~q t > l I 9’

and the previous relationship yields

t
V ~ = Cexp J dr — Cexp (2iO(r)),
o
where
e{T) =
hf{qt~l)dT (3.12.10)
0

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
£ + C exp (2 iO(r)) 1
z — exp
— C£ exp (2iQ(r))

Multiplying the numerator and the denominator by the factor


t

exp 0(T) usdt


3.12 Darboux’s equation 145

and introducing the


notation
a

where
t t

o o
(3.12.12)
we arrive at the expression for the general solution of the
Darboux differential equation in the form of a rational
transformation

where the factors —d, 5, 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
mz' +
n pz' +
q
and require that m, n,p, q meet the conditions
md — pb = 6, —nd + qb = /?, rrido — pbo = <50, —ndo + qbo = /30,
me — pa — 7, —nc + qa = a, raco — poo = 70, — nco + qoo = c^o-

Eliminating for example n and q from the equations


containing a, ao, /?0 on the right hand side, we arrive at the
relationship which determines a
a c —a ao CQ — ao
= 0.
Po —bo
Thus we obtain expressions for Cayley-Klein’s parameters in the form
a = a0 (ad0 — cb0) + /30 (ca0 - ac0), '
P = a0 (bd0 - dbo) + /30 (da0 - bc0), ^ (3.12.14)
7 = 70 (ad0 - cb0) + (ca0 - ac0),
6 = 7o (bdo — dbo) + <$o (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 ~ — K\ — [exp (—i©i) + ££0


AA0
ex
bd0 - db0 = Pi ^ P(*©2)] , [fo exp (*0i) - £
AA^ (3.12.15)
cao — aco = x2 =: exp (-*©2)], [£exp(z@2) —
AA^
dao — bco = p2 £0 exp(—i@i)] : [exp (z© 1)
AAo
where— subscript 0 denotes+ ££0values
ex
P ( ^©2)] ,
relating to the initial values
of the particular solution £ and Ao — yjl + toto-
Thus, the problem of determining the position of a rigid body
rotating 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.
► (3.13.1)
Equations of rotation of a body about a fixed point, known
as Euler’s
equations, are easy to integrate. We
have
Auj\ + (C — A) L02^3 = mh
AC02 — (C — A) (J3U1 =
m 2,
= ra3.
CCJZ
The latter equation yields
3.13 An example. The position of a self-excited rigid body 147

whose first integral is given by


t t
+. tvo 1 +f, . x .C-A
(m2 f+ im\) exp I i———
1
0
cd2 + iu) i = ^2
2 cus dt dt

t
.C-A
exp / cd3 dt dt. (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 m\ = m2 = m3 = 0 we have
C-A „ —
ex —
(J3 = UJ3, (jj T iwi = (^2 + ^1) P ( ^
2

= id0 exp I —1 ~C - A~dd^t T 2£ (3.13.4)

where a;0 and —2s denote the absolute value and the argument
of u;!] By means of formulae (12.6) and (12.7) we obtain
2r
t= q= exp (—2i [r cot A + e]), (3.13.5
Ldu )
where A is given by
cot A = ——pr.0
Aid
(3.13.6
)
The Rikatti equation (12.8) takes the form
= exp (2i [r cot A + e]) + exp (—2i [r cot A + e]) (2,
(3.13.7
and its particular solution is sought in the form
£ = Ai exp (2i [r cot A + e]),
where A is a constant which should be determined from the
quadratic
equation
A2 - 2Acot A - 1 = 0.

(3.13.8)

The roots are A\ — cot ^ and A\ = — tan~. Taking the first root we
148 3. Theory of finite rotations of rigid bodies

and calculation with the help of formulae (12.10)-(12.12) yields 0 (r)

= r cot 0i = r ^cot ^ ^ cot A\ , 02 = r fcot ^ ^ cot A

Then by virtue of eq. (12.15) we find


A A
= sim — exp ( ir — cot A ) I exp ( —ir cot — ) +
>c\
C
A
cot^ — exp I ir tan - P2 A x '
— cot A — 2s
O
Pi = - sin A exp ( —ir exp ( ir cot — J —
-X<1
exp ( — ir tan —
(3.13.1
0)
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
and moreover
sin — exp 2
1
a = cos — exp I i—-— (3 = i = Pi-
Z \ Zd
(3.13.11
What remains is to separate the real and the imaginary ) part of
these expressions.
When a motion differs from regular precession the problem
becomes much more difficult. Consider, for instance, the case
of m\ = m2 = 0, but 777-3 7^ 0- Keeping the above notation for the
independent variable r and constants ct;0,£,cotA and
introducing the relative angular velocity 0 = 073/073 we rewrite
expression (3) in the form r
0U2 + iwi — exp I —2i 1 — — ) cot A / Odr + e (3.13.12)
' 0
Instead of the Rikatti equation we turn to the system of
linear equations (11.4) and introduce new unknown variables k
and r by means of the following relationships

k = exp I i ^1 - ^ cot A J Odr 3


0

r= pexp —1 ^1 — cot A J Odr +£ (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' = ikO cot A -hr, r' = —irQ cot A — k,

(3.13.14)
where a prime implies differentiation with respect to r.
The case 6 = 1 corresponds to the above case of the regular
precession. Obviously, the solution of the obtained system of
linear differential equations leads to relationships ( 10) under
the initial condition >c(0) = 1 and p ( 0) = 0.
The system (14) of two differential equations of first order
can be reduced to a single equation of second order
k" -h k (l -h 62 cot2 A — i6' cot A) = 0.
(3.13.15)
In particular, when the moment of forces about axis Ozf is not
time- dependent, variable 6 depends linearly on time

» = 1 + ^=1 + l, (3.13.16,
<9r,

to
_1

A. I. Lurie, Analytical Mechanics © Springer-


Verlag Berlin Heidelberg 2002
152 4. Basic dynamic quantities

We now introduce the notation that


_ ^ drt drt _ ^ / dxi dxt dyl dy% dzl dzt
Ask Aks — "kTii ^ ' o — / J (o o T r* o To o J >
i=i <T/s \uqs uqk dqs dqk oqs dqk )
R _v'm. ^ i . ^ E i _ v ' m (dXidXi . Qy*dy* . dz
->
s
— * 9f;s dt ^ * \ 9qs dt dqs dt dqs dt J ’
1 A dr* 2
1 - A / dxi \ 2
/ % \ 2
/ dzi \
T o = m
^ -* "SHUO U) UJ+ +

(4.1.2)
Adopting notation (1.2.1) and assuming that
rrii = m2 = m3 = mi, 777,4 = ms = me = m2

etc. we can also write


y'fh?k?kl B - T
Ask — A, fcs ^ dqs dqk
2=1 ’ ’ T°~ ^
Ss
2 — 1
<9gs dt m
(4.1.3)
These values depends upon the generalised coordinates and
time £, however there may exist cases when variable t, being
incorporated explicitly in expressions (1.2.9) for Cartesian
coordinates in terms of generalised coordinates, does not
appear in Ask, Bs and To.
The kinetic energy can be represented as the sum of three
T = T2 + Ti + TQ. (4.1.4)
The first term, i.e. T2, contains the components which are
quadratic with respect to the generalised velocities
nn
T
* = 2EE ASkqsqk- (4.1.5)
■S — 1 fc = l
The second term is linear with respect to the generalised
velocities
n
Ti='S^Bsqs. (4.1.6)
S= 1

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
r*) in terms of the generalised coordinates. Then it follows
from eq. (2) that
Bs =0, T0 = 0, (4.1.7)
An A12 • . . Ain
A21 A22 • • • A2n
(4.1.9)
An 1 An2 ••• A
n
nn
An 1 A
A
••• nn
An • • • Ai^n—1

1,1 • • • An—l^n— 1

A12
A22 >0, Ai = An

for any values of the generalised coordinates in the domain of


definition,
see (A.3.25).
The Sylvester inequalities are as follows
Aii ... Ain
An = \A\ = >0,

An-i >0, (4.1.10)

A2 = >0.

The first inequality expresses also that matrix A is non-


singular, the second inequality says that the matrix of
coefficients of the quadratic form, corresponding to the kinetic
energy of the system subjected to the constraint qn = const, 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 X2 is the kinetic energy of the virtual
dr, 2
YlmiV'i (4.1.11)
i=1 1=1
dt
154 4. Basic dynamic quantities

In what follows we use the Euler theorem on homogeneous


A function ip (x 1, £2, . . . ? xn) of n variables satisfying the
functions.
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

(4.1.14)

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 quasivelocities as being given by
homogeneous forms of the generalised velocities. Then,
using the expressions for the velocities
(4.1.15)

we obtain
nn
(4.1.16)
s=l k= 1
where similar to eq. (2)

(4.1.17)
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

Here we obtain

(4.1.19)

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.
dvi
In the case of stationary constraints the terms with vanish.
When the constraints are non-stationary and the quasi-
velocities are introduced by homogeneous expressions ( 1.5.1)
the terms with coefficients 5s,n+i vanish.

4.1 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
Ps TTT" (s = 1, • • • ,
=
n) , dqs (4.2.1)
156 4. Basic dynamic quantities

see also (10.2.1). In the case of stationary constraints the


generalised momenta are homogeneous linear functions of
the generalised velocities
n
=
Ps ^ ^ AgkQk •>

(4.2.2)
k=1

and in the case of non-stationary constraints


n
Ps
=
^ ^ AskQk “I” Bs.

(4.2.3)
k=i
These equations are solvable for the generalised velocities
since, as pointed out above, matrix A is non-singular and thus
the inverse matrix A~1 exists. By virtue of (A.2.32) the kinetic
energy has the following matrix form
T =-qf Aq + qf B + TQ, (4.2.4)

p = Aq + B, (4.2.5)

which enables an equation for the generalised velocities in


terms of the momenta

q = A 1
(p- B).

(4.2.6)
For stationary constraints we have
T
= \^'M, P = M, 4 = A~xp, (4.2.7)

which leads to the following result

qf = pf (A~1)' = p'A-1,

(4.2.8)
as matrices A and A are symmetric. The expression for the
-1

kinetic energy in terms of momenta, denoted as T', takes the


form
4.2 Associate expression for the kinetic energy 157

For example, for a system with two degrees of


freedom
A-1 = 1 A22 —A 12 —A (4.2.10)
AnAo A? 0 12 An
ill ^22 — ^12
and the associate expression for the kinetic energy is given by

T' = 2AnA22 - A\ ~~ + A11P2) •


(4.2.11)
We notice also the bilinear representation for the kinetic
energy

T = ^Y^qsPs = ^q'p = \p'q• (4.2.12)

f= - T = 2T2 + Ti - T = T2 - T0
(4.2.13)
S= 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 T
f' = q'p -T
= (p' - B') A~lp B') J4 [p-B)~ (p' - B') A~XB - T0
-1

or
P = ± ( p ' - B ' ) A - 1( p - B ) - T 0 . (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
UJ = a <7, q — a~ uo — boo.
l
(4.2.15)

Inserting the latter into eq. (4) we have


T = iJb'Abw + to'b'B + T0 = ^wfA*u + ui'B* + T0, (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.

The first equality in this equation implies another form of eq.


(1.17) whilst
the second is eq. (1.20) for bs,n+1.
158 4. Basic dynamic quantities

4.2 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 YQ is determined by the equation
rc =
1N
ji^lmiri- t4-3-1)
2=1
Here the mass of the system which is the sum of particle
masses
N

M =

(4.3.2)
2=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
1N
rc = r0 + r'c = r0 + — ^ TO* r',
(4.3.3)
2 = 1

where ro = OO denotes the position vector of the pole O which is


the origin of the axes Ox'y'z' fixed in the body, and v'c = 0(3
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 0° 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 premultiplying 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 components of tensor P in these axes as
Psk, this tensor can be represented as a sum of the nine
dyadics 3 3

(4-3-4)
s=1 k=1
1 0 0
0 1 0
0 0 1

3 3 3

a x p = y ^ y ^ y ^ Pkt&S€Skr^-r^-t' (4.3.11)
8 = 1 k= 1 t= 1
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 UJ with
respect to an inertial coordinate system. Applying the
formulae for differentiation of unit vectors
(2.7.6) , we obtain

p= EE = E E {PM+
\s=lk=l / 8=1k=l
Psk [(<*> X i') \'
k + ia (c*> x ife)] } = P 4 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
Psk = Pks (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=EE Psk&sQ'k (4.3.15)


s— 1 k—1
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 O which is the origin of the
coordinate basis. Let e designate the unit vector of axis OA.
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
where stands for the position vector of the point under
consideration and di denotes the angle between and axis OA.
The moment of inertia about this axis is equal to
N N
=

2=1 2=1 (4.3.16)


Jo A = ^ mihi

The square of the scalar product e • can be cast as follows


(e • r*)2 - (e • r*) (r* • e) = e • r*r* • e.

We can also write


e • E • e = 1, r\ — e • Erf • 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 (4.3.19)
inertia
about the axis N

Jo A = e' E (r*2E “ rir*) ' e


mi
0ii ©1 ©13 \
©2 ©22

0D
©23 1 (4.3.21)

C
5
1 2
©31 ©32 ©33 /

N N N
©12
i>2 m
* (vi + 4), = -E miXiVi , ©13 — -E miZiXi ,
2=1 2=1 2=1
N N N

-E JX
rriiXiyi, 022
Jxy
=E mi (zf ©23 = -E miyiZi ,
Jxz \
Jyx Jy — Jyz I• (4.3.24)
Jzx — Jzy
^J

2= 2=1 2=1
1 N N
^rriiZiXi, e = - ^2miyiZi, 633 = '22 i (4 + Vi)
N
32
m

2=1 2=1 2 =1
cf. (A.2.21). The diagonal components represent the moments of inertia about the axes

0n = Jx, 022 = Jy 5 033 = J z . (4.3.22)


The non-diagonal components of the opposite sign are referred to as the products of
inertia and are denoted as follows

©12 = 021 = J y> X ©23 — ©32 = ~Jyz > ©31 = ©13 = ~Jzx- (4.3.23)
In terms of the introduced notation, the tensor of inertia is written in the form

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 0° about point O we look for the
tensor of inertia 0° about point O'. To this end, it is sufficient to
set in eq. (3.20)
Yi = r0 + r',
where OO' — TQ denotes the position vector of the new origin O'
relative to the previous one and O'Mi = denotes the position
vector of point M' under consideration relative to O'. Since
r
i = ro + r? + 2r0 • r', TiTi = r'r- + r'r0 + r0r• + r0r0, the

following relationship is found


N N N

0° = (Erf - r*ri) + Er0 22


^2 mi ° ’ 22 mir>iE
+ 2r

2=1 2—1 2—1


N \ N N

^2 miri ) r0 - TO Y2 miT'i ~ r°r° mi


( 2=1 / 2—1 2=1
or by eqs. (3.2) and (3.3)
e°= 0°' + M (Erg - rgrg) + 2M Er0 • r'c - i (r’cr0 + r0r'c)
(4.4.1)
This relationship is considerably simplified when point O'
coincides with the centre of inertia. Then r'c — 0 and
0° = Qc+M (Er2c - rcrc),
where rc is the position vector 0(3 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 O 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 O and O' we find from eq. (1) the following formulae
defining the components of the tensor of inertia about point O
in o
terms of its components about
' [(Xg + XQ2 + X Q3) 6ik point O'
- XoiXok]
0 ik
x
(4.4.3)
■Oix'ci + XQ2XQ2 “f 03x'cs)
x

&ik ~ 7^ ( ■ 'TA.

Here #oi, ^02, ^03 denote the coordinates of point O' in the
coordinate system Ox 1X2X3 with the origin at point O and x'ci, x'C2,
x'cs 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
Ofx[ etc. are assumed to be parallel. When O' coincides with the
centre of inertia C of the system then x'Ci — 0 and XQI — xci•
Equation (3) takes the form
®?k — ®?k + M [{x2ci + X2C2 +x
2
C2 ) 6ik - xaxck\ • (4.4.4)
For example,
On = ©n + M ( XQ2 + xcz) ? =
0f2
©?2 — Mxc\Xc2•
(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 O with respect
to two coordinate systems and Ox 1X2X3 having the same origin.
Let a denote the matrix of direction cosines which brings the
x= £ = ocx,
(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
0 = ^m(^-4O- (4-
4.7)
The elements of this matrix are related to axes 0£ 1£2£3. The
components of the tensor of inertia about the same point O
with respect to the axes Ox 1X2X3 produces a matrix denoted
by 0* which is given by
0* = ^ m {Ex'x — xxf)

(4.4.8)
or by virtue of eq. (6)
0* = ^ m {Ega'aS, - a&'a') = ^ m (E£ £ - V). r
and moreover
0* ^2 ™ (aE£'£a' - a&'a') = a [^Tm (££'£ - ££') a
as matrices a and a' coincide for all components of the sum.
We then have that
0* = a0a', 0 = a'0*a. (4.4.9)
©11 ©12 ©13

©12 ©22 ©23


> 0. (4.4.1
©13 ©23 ©33 2)
©I 0 0
*
II

©2
0 0

0 0
©3

An expanded expression for the transformation of the


products of inertia is given by
0 1* ©11^11^21 + ©22042^22 + ©33^13^23 + ©12 (O4I<422 + 042<^2l) +023

2 (042£*23 + 043^22) + ©31 (043^21 + alla23) • (4.4.13)

4.5 The principal axes of inertia

It follows from eq. (4.10) that

aQ = 0*o or £©>** (4.5.1)


k=1 k=1

It is known that the directions of axes Ox 1X2X3 (in other words


matrix a) can be defined such that the inertia matrix about point
O is diagonal

0 (4.5.2)
0n -e* ©12 ©13

©12 ©22 ~ ©* ©23

0
CO
CO
©23
0

I-'
CO

0\ 0\
CO
0\ 01 0i
I-'
CO

CO

0\
CO

(4.5.6)

implying that the determinant of the following matrix


/(©*) = 0-^0* (4.5.7)
vanishes. It can be proved that all three roots O s of this cubic
equation are real-valued since matrix 0 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 determinant of determinant (6) for any
5 = 1,2,3. Let the minor determinant of the element 033 — 0* is
non-zero, i.e. ©ii — ©.* ©12
©12 ©22 — 7^ 0,
©g
where 0* is one of the roots. Under this condition the unknown
variables P\ and can be expressed in terms of ^ 3, the latter
being determined from the normalisation condition (5). For 5 =
1,2,3 we obtain three directions and form the matrix \\/3k\\

(4.5.8)
166 4. Basic dynamic quantities

In order to prove the mutual orthogonality of these directions


we notice that due to eq. (4)
3

Y/f3Sk®kt = &*sf3l (* = 1,2,3),


k= 1
which yields

t=l k= 1 t=1

or by means of interchanging superscripts

t= 1 k= 1 t= 1

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 Qkt = Qtk- For this reason

(©:-e;)^M = o,
t= 1

and because of 0* ^ 0* we have


3

E« = 0 (r?s).
t= 1

The latter indicates that the directions corresponding to the


principal axes are orthogonal. Matrix ( 8) is the required
rotation matrix a.
b) If equation (6) has two equal roots = 02, then the
deficiency of matrix (7) for 0* = OJ = 02 is equal to two and
there is a minor determinant of the first order, say ©n —
= 0n — 02, which is not equal
to zero. For any triple of values (5\ and (3^we ^ave one equation (4)
for t = 1 and one equation (5). These values are also related
P\Pi+P12Pl + Pll%=0,
which expresses the orthogonality of the directions
corresponding to the first and the second axes of inertia. The
direction of the third axis is determined 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 (5\ and indicates that one of the principal
axes (the first or the second) may have an arbitrary direction
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 undetermined, 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 O
are given by the unit vectors i*, the dyadic representation of
(4.5.9)
The orthonormalised trihedron i|, i%, is determined uniquely
when the
principal moments of inertia are different If 0* = 0* ^ 0* then the
trihedron is determined to be a rotation about ig. In this case,
due to eq.
(3.8) the inertia tensor can be cast in the form

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
0° = 0JE,

(4.5.11)
with any orthogonal trihedron being taken as the trihedron of

4.6 Inertia ellipsoid


Construction of the inertia ellipsoid serves as an auxiliary
means to illustrate the concept of the inertia tensor at point O.
Let us consider a line segment of length p

1
P= y/ Jo A
(4.6.1)

along the axis OA whose direction is given by the unit vector e.


The coordinates of point K which lies at the end of this
segment are
a P (4.6.2)
y/l^A y = y/JoA ’ y = V Jo ’ A
168 4. Basic dynamic quantities

where a, /?, 7 designate the direction cosines of the angles


between vector e and the axes Oxyz. These are the direction
cosines 011,072, <^13 in eq. (4.11) provided that 0^ is replaced
by Jo A- Performing this replacement and inserting a, /?,7 from
eq. (2) we obtain, after some simplifications, that
011X2 + &22V2 + ©33 z2 + 2©12 xy + 2©23 yz + 2 Q31ZX = 1.
(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
©1X*2 + ©2^*2 + ©3 2*2 = 1. (4.6.4)
Therefore, the axes of the inertia ellipsoid coincide with the
principal axes of inertia at point O, the length of each axis,
due to eq. (1), being given by
1
(4.6.5)

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
^mx*2 < ^mi/*2 < y^mz*2,

then
©*>©*>©* 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 ©3 = 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
©I <©2 + ©3> i-e. +
(4.6.7)
4.6 Inertia ellipsoid169

FIGURE 4.1.
We assume that 0^0 7and^ introduce into
consideration
£1 a? ©2 a
2
©3 (4.6.8)
“0*’ ^2 0*‘
£2 = =
Inequalities (6) yields

£1 < 1, — < 1.

In the plane £1, £2 they define a domain which is the interior of


the rectangular 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 Ox* 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
£1 + £2 ^ 1.
One can see that the inertia ellipsoids correspond to the
hatched domain
ACB 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
£1 = £2 = 0.5, i.e. ©1 = 2©2 = 2©3. It is worth noting that the non-
central i.e. ©^ = 0.707©i, ©3 = 0.293©!.
Kovalevskaya ellipsoid for which ©* = ©*= 2©3 is situated in at
point
170 4. Basic dynamic quantities

This is the inertia ellipsoid which deviates most from the


ellipsoid of revolution.
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 VQ — 0

v^wxr'.

T
1N iV
1(4.7.1)
= ■z'52rniVi-Vi = -^2m i(u;xr'i)-(u>xr i) =
,

1= i=l
1 N
N
x (" x r')] = x mi Lo2r,/2- («• ri) •
i=1 i=l
This equation is transformed using eq. (3.16)
w2rf - (w • r') (w • r•) = u> • (rf E - r'r') • u>.
Using the definition of a tensor (3.20), we yield the following
formula
r=lfa>-0°-«,
(4.7.2)
z
where 0° denotes the tensor of inertia of the rigid body at
point O. The kinetic energy is represented by a quadratic form
of projections UJ\,002,^3 of the angular velocity vector u with the
coefficients defined by the symmetric tensor 0°. The matrix
form is as follows
T = lfa/e°fa;, (4.7.3)
z
where uo and u/ are a column-matrix and a row-matrix of
projections of UJ on the taken coordinate axes and 0° is the
matrix of inertia.
Particularly, if axes Ox'y'z' fixed in the body are taken as the
coordinate axes, then 0^ are constants and the above form
has constant coefficients. An expanded form of eq. ( 2) or (3)
is
4.7 The kinetic energy of rigid body 171

This equation simplifies when the principal axes of inertia at


point O are used to
+ + (4-
7.5)
The above formulae express the kinetic energy in terms of
the quasivelocities. When the generalised velocities ip, ip are
used the expression for the kinetic energy becomes
cumbersome even for the case in which the principal axes are
taken
T = ^ { [(0J sin2 ip + 02 cos2 ip) sin2 # + ©3 cos2 #] ip2 +
(0i cos2 ip + ©2 sin2 ip) $2 + ©3p2 + 2&^ipp cos $ +
2 (©i — © 2) sin cosy? sin .
(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
and thus
Vi-Vi =v0 + 2 (v0 x u) • r• + (a; x r-) • {UJ x r •).
By using eq. (2) and definitions (3.2) and (3.1) we obtain
T = - [Mv0 + 2M (v0 x u>) • Y' + u> • 0° • cj] .
C

(4.7.7)
Here v0 is the velocity of the pole O, r'c denotes the position
vector 0(3 of the centre of inertia of the body in the axes with
the origin at the pole O, and 0° 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 ^ 01,^02,^03
of velocity of the pole O 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 (^OI + ^02 + ^03) + 2M [(v^s — ^03^2) x'c+
{VQ^LUI — foi^3) Uc 4” (^01^2 —
^02^l) %c\ 4" ©1^1 T ©2^2 4” ®3^1} •
(4.7.8
)
The above expression is drastically simplified when the
centre of inertia is taken as the pole. Then x'c = y'c = z'c = 0 and
172 4. Basic dynamic quantities

Another important particular case is that of rotation of a rigid


body about a fixed point O, when the inertia ellipsoid at point
O is an ellipsoid of revolution with the axis defined by the unit
vector e. By virtue of eqs.
(5.10) and (2) we obtain
0X + (0|-0^ ) ( u , - e )2 (4.7.10)

where • e is the projection of the angular velocity vector on


LO

the axis e, and ©3 and ©i 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 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 = rnjVj = y; mA.
(4.8.1)
2=1 2=1
By virtue of eq. (3.1)
1N
rc = vc = b> m<
(4.8.2)
= 2 1

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 O, i.e. the geometric sum of the moments of
momentum about this point is given by
_N
K° = x Vj.
(4.8.4)
Q =Mvc = M (v + x + Mv'c = Qe + Qr,
0 (JO r'c) (4.8.5)
4.8 Principal momentum and principal angular momentum of a rigid body
173
where r'c=0(3 denotes the position vector of the centre of inertia
with respect to the above axes, WQ and v'c are the absolute and
the relative velocities of the centre of inertia, respectively, and
Qe = M (v + w x rrc), Qr = Mv'c
0

(4.8.6)
designate the vectors of the translational and relative momenta ,
respectively.
The expression for the angular momentum can be
_N
K = 53
5
(r0 + r') x (v0 + x r• + v')
2 = 1
N N

= Mr o x ( v 0 + UJ x ry+Mr'cxv0x + X r')+^ m ^ r - x v - .
2=1 2=1
We have
N N
miT
53 'i X (W X T'i) = 53 ~ Tiri ' “)
2=1 2=1

= 53TOi= (Eri2 ~ r i r0 = ®° ■ “,
2 1

where 0° denotes the inertia tensor at point O. Due to eq. (4)

53m*ri xv'=Kr°

(4.8.7)
2=1
represents the relative angular momentum about point O and we arrive
at the equality
K° = r 0 x Qe + Mv'c x v0 + 0° • u + K°.
(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
We then have
Q= M (v0 + u> X rc), (4.8.9)

K° = ro x Q + Mr'c x VQ + 0° ■ u>, (4.8.10)


174 4. Basic dynamic quantities

where 0° is the inertia tensor of the rigid body at the pole O of


the axes fixed in the body. In particular, if the centre of inertia
is taken as this pole, then r'c = 0, r0 = and
Q= Mvc, K ° = r c x M v c + 0c - w. (4.8.11)

When the rigid body has a fixed point O and this point is
assumed to coincide with O then ro = 0, Vo = 0 and
Q = Mu x r'c, K° = 0° • u>.

The latter expression, being projected on the axes Ox'y'z'


fixed in the body, takes the form
K\ = ©11^1 + ©12^2 + ©13^35
K2 = ©21^1 + ©22^2 + ©23^35 > (4.8.13)
= ©31^1 + ©32^2 + ©33^35 ,

since the products of inertia are equal to Qst (s ^ t) with the


opposite sign. When the principal axes of inertia at point O are
taken as these axes then
K° = eXuju K® = ©2^2, K° = e*3Lj3. (4.8.14)
Returning to Sec. 4.3 it is worthwhile mentioning the special
case of the body rotating about a fixed point O 'in which the
inertia ellipsoid is an ellipsoid of revolution. Then
K° = (w - u) • ee) + ©£u; • 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 u) 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

= VQTW x r'+v',
where Vo and LO denote velocity of the pole O and the angular
velocity of the moving axes, respectively, is the velocity
relative to these axes and
4.9 The kinetic energy of a system under relative motion 175

v* is the absolute velocity of the particle. Using the well-known


equality
(a x b) • c = a • (b x c) we obtain
vf = vl + 2 (v0
r' + (u x r') • (u> x r') +
X u>) ■
• v' + 2u> • (r- x v') + vf. 2v0
By means of eqs. (7.7), (8.6) and (8.7) we find
T= ^2 [Mvo + 2M (v0 x u) -r'c + u • 0° • a;] +
N
v0 1 (4.9.1)
Q r + u > - K ° + -^TOiuf.
i=1
The first line of this 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 0°
at point O does not remain constant under a motion (which
would be the case of a rigid body fixed with respect to these
axes). The term
1N ,/2 (4.9.2)
ME miV,
i—1
represents the kinetic
energy of relative motion. Now we can write
T = Te + v0 • Qr + + Tr.
(4.9.3)
Describing the position of the system’s particles relative to the
moving axes by means of the generalised coordinates q\, q2, . . . ,
qn and assuming r
i =ri ,Qn), (4.9.4
)
we have
N N
dr'
Qr = ^rriiVi =
(4.9.5
i=1 s=l i=1
N N
dr'
K
r = ^2miri X = J2^12miri X df/ (4.9.6)
i=1 S =1 2=1

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 (4.9.7)
T\ = vo • Qr + CJ • Bsqs,
s= 1
176 4. Basic dynamic quantities

where

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 qs and is understood
as TQ 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)
i=1

which is termed, by analogy with the kinetic energy, the energy


of accelerations. There is no need to derive an exact expression
for S since Appell’s equations contains only derivatives with
respect to the generalised accelerations qs. The terms which
do not depend upon qs 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
N
= Emi i 2=1 ( d2n dr i
dqs \dqsdqk dqr

dArs II ( d2n dr,


dqk ^ &Qk &Qr dqs

Subtracting the first expression from the sum of the second


and third expressions we have

dri 1 f
N
d2n
rriidqsdq
E dArs dAkr dAsk (4.10.2)
_
k 2 V 0qk
dqr
The expression on the right-hand side is referred to as a
Christoff el’s symbol of the first kind for matrix |Asfc||. The notation is [k, s: r] =
[s, k; r\ which is known as Christoff el’s square brackets, see (B.4.14)
1 f (4.10.3)
[s,fc;r] dAsr ^ dAkr dAsk
2 V
Now we obtain
^nn nnn
s- = | E E AskQsQk EEE [s,k;r]qsqkqr. (4.10.4)
s= 1 k=1 r= 1 s=l k=1
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 s and k and over r in the second sum is from
1 to n (as qn+1 = 0) whereas the summation in the second sum
over sn+and
1n+1k is from 1 to n + n
1. n
The result is
E E [s, k; r] qsqk E E [s, k; r] qsqk +
s=l k— 1 s=l k=1
n
2^[s,n+ l;r]<7s + [n+ l,n + l;r].
S= 1
2T0, An+l,r

n.
E 'dAsr . dt +
drt drt
s=lX t=i
mi
dqr dt ~Br'
E

dT0 (4.10.5)
dt
1
1^

nn n
Qr,
dqs dqr

^ 7i 7i nnn
*' - ; E E AskQsQk T E E E [s, A; r] + (4.10.6)
s=l s=l fe=l r=l
k=1
71 71 / dAsr dBr dBs \ .. . / dBr dT§ \ ..
SS (-9T + w, - W ) M - + E (-ar - w) *■

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 LUs. 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
dr, d2i
i Eisr^+EE ~UJk^S’
s=l k=1 d-Kkdir,
w

s—1

In view of this we can write


S
nn N 0 0

= 5EE“>"-E”<9^'9j; + (4.10.7)
s=l k—1 2=1
I nVn
n I VI V IV dviQ( d2Vi d2r, \
\ ^ \ ^ Or i dirkdirs ditsditk J
+
I
r —1 s=l k—1 I
dn d2ri 2
It should be stressed that ——^ , see eq. (1.9.5). A
further
OTTkdTTs U7T sU7Tk
calculation is carried out fully analogous to the above. By
4.10 Energy of accelerations 179

(1.17) we have
dA*r _ -A / dri_ dr£ d2r, \
d2rj '
ditk \d7Ts dirkdirr 9K r dirkd-Kg)’
9A
*k_S^m (dri d 2r
‘ >- | dr
i d*r* ) „ (4.10.8)
dlTs 4^ 1 \dlT r d'KgdlTk dlTk d'Ksd'Kr) ’
2=1 X

dA*ks y> / drt2


d rj drt d2rt \
9K r d'KrdiTs 9K s dKrdKk) ’ y

Recalling the rule (1.5.17) of ”differentiating with respect to


the quasicoordinate” and introducing the generalised
Christoffel symbols we have
1 / dA*sr dAjr dA*sr
dA*k d([ +
2 \ d7Tk dlT s m
dA* dA* \
b m s - bmr—^ = [s, k\ r]n = [fc, s; r]n .

(4.10.9)
dqm dqrn )
Subtracting now the third line in eq.( 8) from the sum of the
■ , , _ 1s if d2rj
’ ’2^—'m'ldTrrdr\dTrkdn s dnsdnk
d2rj dri
2=1
X
rrii 9K
S
/ d2rj________d2r, \ 1 y^ drj_
l d'K^d'Kr d'Krd'Kk ) 2 ^ dlTfc d2 rt d2rj \
x
' 2=1 d'Ksd'Kr dKrdKsJ
(4.10.10)
The difference of the ’’second derivatives”
d2n d2rt
dKf-dKr d'Krd'Kk
can be expressed by eq. (1.9.5) in terms of the ’’first
derivatives”
<92rj_____<92r, _ AA t dr^
dKkdKr dKrdKk “ ^ dKl ’

d2ft______d2rt
_ AA j drj_
dKgdKr dKrdKs “ 7rs dKl '
We obtain
ly 0T£ / f)2r,____________<92rz \ _ lA |2 y dri drj
2 ~^mi 9K s \ dKkdKr dKrdKk ) ^ ^^^

=
z
Z=1
180 4. Basic dynamic quantities

and by analogy

}_ 1 d2Ti \ _ 1 rs kl i .*
2^-' dirk \d7rsd7rr d7Trd7T s) 2^-^^ 2=1 N7
2=1

Inserting this into eq. (10) and recalling that= -l^lrkkr y 1


l ields N

2EW^+^)-
2=1 Z

1^ dn dr <9 r;
2 2
r
(4.10.11)
= M; L +
9.z dlV b dlT s dTT.dTTi
2=1 2mi

Equation (7) now takes the form

r = l±± A-sk^s&k + 2 ^ v ^ I ^ ^ uj cu cuk / [5, A:; r]^ + ^ ^ 7sr^-ki


r s
z
s=i fc=i z
r=l s=l fc=l l 2=1
(4.10.12)
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 O as the coordinate axes we obtain due to eq.
(7.5)
for s ^ k, A*s = 0S = const.
A*sk = @sk =0
Hence, all expressions [s^k'^r]^ — 0. Using eq. (2.10.3) we
obtain
5* = i (©xw? + 02^1 + 03w§)
(4.10.13)
-hc5i(J2^3 (©3 — ©2) + 652^3^1 (©1 — ©3) + UJ3<J0\(J02 (©2 — ©l) •

Let us notice that the expression for S* in terms of the

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 O
which is taken as the origin of the basis axes fixed in the
body. We have
Wj = CJ x r• -b u? x (u? x r').
4.11 Energy of accelerations of a rigid body 181

Retaining only the terms with Co we obtain


N N
S* = - ] > > ( * x
r ' ) • (u> x r ' ) + ^ m , ( w x r ' ) • [ w x ( u > x r ' ) ] .
2=1 2=1
(4.11.
1)
The first term in the latter equation differs from the equation
for the kinetic energy T in that Co occupies the place of oo.
Then, due to eq. (7.2)
1N1
2 ^2mi x r
'i ) • i) = 2^ ■ ®°-&-
x r

(4.11.2)
2=1
The component of the second sum in eq. (1) is transformed
as follows
(d> x r') • [u> x (u> x r')] = (w x r •) • (ww • r- - to r') 2

= ( w x w ) -r'u> • r' = (w x o>) • [r- x (u> x r ■ ) ] ,


where, while manipulating, we took into account that
(Co x r•) • r' = 0, (Co x r') • uo = (oo x Co) • r', ( w x w ) * w = 0.

The second sum in eq. (1) is then given by


N N
^ ra; (u> x r') • [w x (w x r •)] = ( w x w ) ' ^ m; r' x (u> x r ' )
2=1 21 =

= (Co x oo) • K°,


(4.11.3)
where K°, due to eq. (8.4), is the principal angular moment of
the rigid body about the fixed point O. Using expression (8.12)
for K° we arrive at the equality
s* = u> + db ■ (u> x K°j = • 0° • d> + u ■ (w x Q°-u).
■ e° ■
(4.11.
4)
The last term also can be written in the form (Co x oo) • 0° • oo
which is a bilinear form of projections of the vectors Co x oo
and oo with coefficients defined by tensor 0°.
where CJ is a row-column of projections of vector cl;, a; is a
column-matrix of
vector co, and Co denotes the skew-symmetric matrix
accompanying vector
182 4. Basic dynamic quantities

An expression for S* in terms of projections of the vectors LO


and CJ and the components of tensor 0° is easily obtained
using eqs. (4) and (5)

S* = — (©lld;^ T ©22^2 4~ @33^3 4“ 2©i2<^lC^2 4" 2023^2^3


4"
4- (UJ2^3 ~ ^2 ^3 ) (011^1 4- ©12^2 4- ©13 ^3 ) 4-
(OJSUI — UJ3UJ1) (©21^1 4- ©22^2 4- ©23 ^3 ) 4-

(Cj\(jJ2 ~ W1W2) (©31^1 4- ©32^2 4- ©33 ^3 ) •

(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=w0 + wxr-+o;x(a;xrJ)

(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 w o and £j
replace and , respectively. With this in view we obtain the
VQ CJ

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
^mjW0 • [<jj x (u; x r')] = Mw0 • [LJ ( r^)] X CJ X

i—1
= M (wo x u) • (UJ x r^).
Thus, in the general case of motion of a rigid body
S* = ^MWQ + M (w 0 x cl;) • r c
f
+ M (wo x CJ) • (u> x r'c)
4-^CJ • ©° • CJ + (cl; x u>) • ©° • u>.

(4.11.8)
A considerable simplification is achieved when the origin of
w0 = Vo =Vo X v0
4.12 Example calculations of the kinetic energy for multi-body systems 183

the projections of the vector VQ on the axes fixed in the body


are equal to time-derivative of the projections of vector . This VQ

enables us to
**
replace in eq. (8) WQ by VQ +2 v0 • (a; x v0) and w0 x UJ by v0 x UJ
and to retain only the terms containing the quasi-
accelerations, i.e. vectors v0 and d>. Simple manipulation
yields
1 *
S* = -M VQ +M (V0 + x r^) • ^v0 +

(4.11.10)

M ^v0 +u> x vo) • (il> x r^) + ■ 0° • CJ+ (CJ X U>) • 0° •

u;.

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
u>2 = dii + (3i2 = i[a cos (3 + (3i'2 + i3a sin /3,
where ii is the unit vector of the axis of rotation of the outer
gimbal, and i's 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 = \ (^A2a2 cos2 (3 + B2(32 + C2a2 sin2 f3^j .
In order to derive an expression for the kinetic energy of the
rotor we write its angular velocity vector in the form
= u}2 H- i'3(f = ijdcosP + (3i'2 + i3 (d sin/? + < £ ).
u)3
Using expression (5.10) for the inertia tensor at point O, we
obtain
T3 = -u>3 • [A3E+ (C3 — A3) igig] • U>3
= 2^d2 cos2 f3 + (3 ^ + — C3 (cp + d sin/?)2 .
184 4. Basic dynamic quantities

The kinetic energy is thus equal to

T= ^ (Ai + A2 COS2 /? + C2 sin2 f3 + A3 cos2 0) a2


-\--(B2-\-AZ)(3 + -Cs ((f + dsin/3) .
(4.12.1)
4-12.2 A shell carrying flywheels
A rigid shell rotating about a fixed point O is considered.
Denoting the principal moments of inertia at point O as A, B, C
and the angular velocity vector as LO we have due to eq. (7.5)

To — - (Acu2 + BLU2 + CUJ2).


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 e* and is assumed to have a
constant direction with respect to the shell described by the
direction cosines of the angles referring to the principal
axes Oxyz of the
shell. The position vector of the centre of inertia Si of the
flywheel is denoted by r* = OS'*. Finally, (p^i stands for the
angular velocity vector of the flywheel relative to the shell and
the absolute angular velocity of the flywheel is given by
oji = v + ei<pi.

Assuming that the equatorial moments of inertia are equal (Ai —


Bi) and denoting the axial moment of inertia by Ci, we obtain
with the help of eqs.
(7.9) and (7.10) that
2Ti = TO* |u> x r*|2 + (w + e*£*) • [A*E+ (C* - A*) e*e*] • (w +
e*v?*)
= rriiU>- 2(Er* • r* - r*r*) • u>+A* - {u- e*)2 + C* (u> • e* +
OJ2

<^>*) .
(4.12.
2)
The kinetic energy is obtained by addition of these
expressions for all
(4.12.3)
flywheels with the kinetic energy of the shell, which results in
the formula
4.12 Example calculations of the kinetic energy for multi-body
185 systems

FIGURE 4.2.

where 0° is the tensor with the components

©?i = A + E[Ai {fi + 72) + m; (7? + z f ) ]


2=1 (4.12.4)
_ n
0?2 = - E {Ai<XiPi + mXiVi)
2=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 O
whose position in the body is given by vector r = OO. The unit
vectors of
the trihedron of the principal axes of inertia at point O are
denoted by
a, b, c and are shown in Fig. 4.2. The unit vectors e i, e 2, e are
fixed in
the shell. The angle p between the plane of the unit vectors b,
c and the
plane e, e± describes the angle of rotation of the flywheel and
A is the angle
between the principal axis of inertia c and the rotation axis.
The position
and hence

2 Te = m \u) x r|2 + 2m (u) x r) • (a; x p) + 2


Ax (w • a)2 + £ ! ( « • b)2 + Ci (w • c) .
an <*12 <*13 sin p — COS(/? 0
9= ot 21 <*22 <*23
, k= cos A cos A sin — sin A
<*3 <*32 <*33 cosAp
sin psin A sin cos A
1 cos p p

The rotation matrix which makes the trihedron of axes Oxyz


coincident with trihedron a, b, c is thus kg.
Let us introduce the row-matrices
li = II 1 0 0 ||, 1'2 = || 0 1 0 ||, T3 = || 0 0 1 ||. (4.12.9)
Then the row-matrices a',b',c' of the projections of the vectors
a, b,c on
axes Oxyz are given by
(4.12.10)
a' = li kg, b'=l'2kg, c'=l'3kg
4.12 Example calculations of the kinetic energy for multi-body

systems 187 and thus


UJ • a =l[kguj, UJ-b =l'2kguj, UJ • c =l'3kguj, (4.12.11)

where denotes the column-matrix of the projections of UJ on


UJ

the mentioned 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 £1,6:2, £3 of vector p on trihedron a, b, c. Then

£ = kgp, p = g'k's

and we have
\UJ x r|2 = UJ • (Er • r — rr) • UJ = uJ (Ex’x — xxf) UJ,

(w x r) • [(UJ + eg>) xp] = UJ • (Ep • r — pr) • (UJ + e<p) >

= UJ' [Ex’g’k’e — g'k'ex') (UJ + gjg'I3).


(4.12.12)
The calculation is not continued, as an extended expression
for the kinetic energy given by eq. ( 8) would be very
cumbersome. However the calculation is somewhat simplified
for the case when = and the rotation axis e is parallel to axis
g E

Oz. Then

UJ • a =l[kuj = 1 sin gj —
UJ UJ 2 cos gj,
UJ • b — cos A 1 cos + a;2 sin — 3 sin A,
=l2kuj (UJ gj gj) UJ

UJ • c =13^0.; = sin A 1 cos + a;2 sin + 3 cos A.


(UJ gj gj) UJ

Provided that the flywheel is only statically unbalanced, i.e.


A = 0 and A\ = Hi, the expression for the kinetic energy takes
the form
T
=V AUJ + BUJ2 +
\ EUJ^ + m \UJ x r| -f- A.\ (UJ'I -h UJ2) C\ (UJ3 -h (p)

+ muj' (Ex'k'e — k'ex') (UJ + <£>13). (4.12.13)

In this case we can assume that £2 = 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 £3 = 0, e = e 1 li and the calculation yields
vnuj (Ex'k'e — k'ex') ( UJ + (/7I3) = me\ {cos gj [( UJ 3 + <p) (UJ 2Z — uj 3 y) —
UJ 1 (uiy - uj 2x)\ - sin gj[uj 2 (uj\y - UJ 2X) - ( UJ 3 + Cp) ( UJ 3 X - UJ \ Z )}} .
188 4. Basic dynamic quantities

FIGURE 4.3.
It is clear that the same result can be obtained by extending
the following vector expression
[(a; x r) x (a; + e(p)] • p
and noticing that the projections of p on axes Oxyz are equal to
e\ sirup, —si cos(p,0.

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 perpendicular 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 O.
The orthogonal trihedron is of the unit vectors is bound to
the outer ring, such that ii and 12 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's with i'2 = 12- The bearings of the gimbals
Kk are fixed to the platform and a/-, hk, ck denote the trihedron
of the orthogonal unit vectors bound to the gimbals Kk,
4.12 Example calculations of the kinetic energy for multi-body systems 189

7fcand the position of the point Ck of intersection of the axes of


rotor Rkand gimbal Kk is given on the platform by the position
vector CCk = r'fc. 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 Cfc.
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 fi + i i d, where a denotes the angle of
rotation about axis ii. Then
T\ = \MXV2 +1 (n + i x d ) • ef • (n + h a ) .
The platform. Its angular velocity is equal to
b)
u p = Q + i1a + i'2P, (4.12.14)
where (3 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 = l [M2v2 + 2M2V- (wp X r*) + • 0 f• uP\ .

The gimbal Kk. Its angular velocity and the velocity of


c)
point Ck are given by u;p + afc7fc and V + up x r'fc, respectively.
The position vector of the centre of inertia Sk is CkSk = In
accordance with eq. (7.7) the
kinetic energy Tk of the gimbal Kk is as follows
TL = l{m'fe|V + u , p X r 'fe|2 + 2m'fe(V + u , p X r ' f e ) -
[(wp + afe7fe) x fffcCfc] + (wp + afc7fe) • 0,Cfc • (u>P + afe7fe)} .
Here 0,Cfc denotes the inertia tensor at point CV Assuming
equality of the equatorial moments of inertia A'k and B'k about
axes and we have
The rotor Rk. Its angular velocity is u;p + a^7k + ck&ki where
d)
Cpk denotes the angular velocity of rotor Rk• As the centre of
inertia coincides with point Ck we obtain
rpH
1
___ 1 m'k |V + up xr'|2 +
k~ 2L
(ujp + afe7fe + ck<pk) ■ 0"
Cfc
• Ft 1^1

where (wp
0 ,/C
*=A£E+(<?£-AZ) cfccfe
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
\MV> = i M2 + ^ {m!k + m'l) V2.
k=1
The term with the factor V x ujp is given by
m
(V X LOp) 'k + mk) 4 + mk£kCk
M2r* + (
k=l fc=i
By properly mounting the balance masses on the platform,
one can determine the centres of inertia of the platform, the
rotors and the gimbals coincident with point C, i.e.
M2r* + ^ ( 'k + k) r'fc = 0-
m m

k= 1
In this case the expression for the term associated with V x up is
as follows
n

(V x u p ) - J 2 ^ k £ k ^ k •
k =1
The terms quadratic with respect to a;p are
1 f 71
-LCP • -I- ( m'k -I- m!k) (Er'fc • rk - r'krk ) 4-

n n ^

[EAfe + (Cfe - Afe) C f c C f e ] + m'fcefc (Er'fe • cfe - r'fecfe) l • u>P.


k=1 fc=l

Here = v4'fc 4- v4'fc' and Ck = Ck-j- C*.'. Let us denote the constant
part of
the tensor in the braces as Q. The latter equation then takes
the form
1 1 n^
-Wp • Q • Wp + -U>p ■ [ ( C k - A k ) C f e C f e + m!kek (Er'fc • ck - r'fecfc)] • u>P.
fc= l
The terms depending upon the angular velocities 7fc of the
gimbals with
respect to the platform are given by
n n ^n
~ mkek (V + upX r'fc) • bfc7fc + £ Akjku)P -ak + - ^ Akj2k.
k—1 n,
k=1 k=1
Finally, the termsE associated
H k W p • C kwith + the angular velocities of
=1' 2C''J’
4.12 Example calculations of the kinetic energy for multi-body
191 systems

FIGURE 4.4.
where Hk = CkCpk 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 = ~ (Mi + M) V2 + - (0 + i i d ) - 0 f - ( Q + ( V x cjp)-^ mnkekck


+ iid)
k=1
+2 Up• |Q + 5Z K ~ )
Ck Ak CfeCfe
+ 'k k (Er'fc • cfe - r'kck)] j • uP
m £

n in
+ m'k£k (v + upx r'fc) • bfe7fe + - ^ Ak^ku>p ■ ak
k=1
Z
fc=l
+
(4-12-
15)

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 Q = 0. The rotation axes of
the gimbals passing through points Ci, C2, C3, C4 are
perpendicular to the plane i2, of the platform and their unit
vectors are
a =a
i 3 = ig, a2 = a4 = —i3.
The gimbals are arranged in pairs so that = y 2 and y3 = y4.
This effect can be reached either by means of a gear train or
192 4. Basic dynamic quantities

an antiparallel 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 Cfc). The position vectors of points Ck are

ri = aii + M 2 + ci3, r'2 = -aii + bi'2 + ci'3,


1*3 = aii - bi2 + ci3, r4 = -aii - bi2 + ci'3.

The masses mf and the inertia moments of all gimbals A', C" and all the rotors (ra",
A", C") are assumed to coincide. The balance masses are absent. The angular velocity
of the platform is given by

u>P = i^dcos/J + i2(3 + i3dsin/J.

Expressions for the relevant vectors are given by

ci = ii sin 7! — i2 cos 7X, C2 = i'i sin 7X + i2 cos 7X,


c3 = -ii cos 73 - i2 sin 73, c4 = ii cos 73 - i'2 sin 73.

Formula (15) for the kinetic energy is considerably simplified and takes the form

T 1©HQ!2 + lfa>p • &2 ■ Up + Imwp • ^ (Er'fc • r'k - r'fcr'fe) • u>P +

2A (a2 + (3 ) + +— (C1 — A) (wp • ck)2 + A (7, + 73) +


k= 1
4
14
ALVP • E ^kik + a;p
’ * (4.12.16)
fc=l k= 1

Assuming knowledge of angles a,/?,7I,73 and taking into account the above
expressions for cjp, rj. and C& we obtain

T — — [©ii + ©2i + 2A + 2Cf + 4m (&2 + c2)] d2 + — [©22 H- 2A + 2(7 z z


2
+4m (c2 + o )] (32 + A (7? + 7§) + 1(7" (v?j + a-f1 -/?) +

(<y?2 + 007 + /?) + (<^3 - a - /?73) + (</?4 + a — ^73) .


(4.12.17)

Here ©n denotes the moment of inertia of the outer ring about its rotation axis, © 21,
©22 denote the moments of inertia of the platform at point C about i[ and i2,
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 Secs. 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 - 1 3 . 1 Rolling sphere
Assuming that the centre of inertia of the sphere coincides
with its geometric centre we have due to (7.7)

T = -M (XQ + 2/Q) + 2® (^i + ^2 + ^3) 5

(4.13.1)
where © denotes the moment of inertia of the sphere about an
axis passing through its centre. For a solid sphere
© = %Ma2.
5
Using expressions (2.10.19) for the quasi-velocities 0)4 and 0)5
T = -M (UJ\ + 0J2) + —n o)3 T 2a£j2&4 ~~ 2auj\Cj^ + <4)4 + 0)5
5
2

(4.13.2)
In order to derive an equation for the energy of
accelerations we turn to eq. (11.9). For the sphere
29
© • (JJ = -Ma
u>.
5
The directions of the vectors © • u; and u? are seen to
coincide, hence
x UJ) • © • CJ = 0
and the latter term in eq. (11.9) vanishes. By virtue of eq.
(2.10.19) we have

x0 =a>4 +a a>2, 2/o =0)5 ~ a


&i •
Therefore
2 -2 , -2 ~ 2 ■ ~ 2 . 2 wc = X0 + ~ 2 , ~
2 Ldl T + 2a 0)20)4 —2a 0)10)5 .
VO =^4 + ^5 +a UJ 2
194 4. Basic dynamic quantities

The term —Co • 0 • Co in eq. (11.9) differs from the


corresponding expression
for the kinetic energy in that Co replaces LO. Hence, while using
projections of the vector on the fixed axes (which is the case)
or on the axes fixed in the body it suffices to replace Co2 by C02
1. . 12 2
- U .Q . U = - -Ma 2 , ~2 , ~2 Icq
T CO2 T
We have
5* = T(u>),
(4.13.3)
which means that in order to construct the expression for the
energy of accelerations it is sufficient to replace the quasi-
velocities Cos in the expression
for the kinetic energy by the quasi-accelerations Cos.

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 r f c = 0. Using eqs. (2.10.25) and (2.10.26) we obtain
VQ = LO2 + LO2 + oo\ = (col — CLLO3)2 + (ic>5 — acoi sin'#)2 + a2col cos2 #
= LO42 + ic?g2 + a2 (1o\ + cof) — 2auo3Uo\ — 2auo\uot> sin#. (4.13.4)
The moments of inertia of an infinitesimal thin ring about
the axis perpendicular to its plane and an arbitrary axis in the
plane are respectively equal to Ma2 and |Ma2, both axes
passing through the centre O. For this reason, the tensor of
inertia at point O is

1 = Ma2 ^^3 + ^nn+ ^n/n/^ •


(4.13.5)
LO - O - LO =Ma
2 2,12,12
^3 + 2^1 + 2^2

Hence
_ n2 3 2 1 2auo3Uo\ — 2auo\uot) sin # + d*2 + uo*2 .
T = -M a I 2^3 + — LO± + —LO.

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,nl5i3
v0 = LU 4I 1 + u?5ni + 00Q I 3 = ~ a (c^n + <*>1 sin$ni - 00 \ cos^ia).
Due to eq. (2.9.7) and taking into account notation (2.10.14)
we obtain ^2 -n, d\3
n ni, ni —— = 0
slu'd
00 2
sin d ’ dt
and thus
V0 = Wo = —CL [(d?3 — OO1OJ2) n+

(001 sind + cosd + LU2LU?.) nix _ (cj1 Cos$ - uo\ sind) i3 7.


V sin d J J
While calculating WQ we drop the terms which do not contain
cl?s, such that
WQ = a2 (CJ\ + 00*2) + 2a2oo2 (001003 - 001603).
(4.13.7)

We consider the derivation of equations for the angular


velocity vector. By virtue of eq. (2.10.13) we have

' 00 = cjin-K^n' +

d? = [001 + 002 (003 002 cot d)] n+ [oo2 - 00\ (003 - 002 cot d)] n' +
^313•
(4.13.
9)
To find the latter term in eq. (11.9) it is sufficient to take into
account only the terms that include 00s. Up to this order of
accuracy we have

Ct? X Ct? = (0O2OO3 ~ 002003) II + (0O1OO3 — OO1OO3) Tl' + (cl?iCt?2 — 0Ji0J2) i . 3

Further on,

0° • 00 — Ma 2
(^LOin + o n' + c^i^
2

(d? x u?) • 0° • OJ = ^Ma2oo3 (oo\oo2 — ooioo2). (4.13.10)


196 4. Basic dynamic quantities

We represent vector UJ in the form

Co =LO + e , e = 0J2 {003 — 00 2 cot d) n—uo\ (cos — 002


cot $) n',
*
where 00 denotes the vector whose projections on the axes of
the ” halfbounded” trihedron are respectively dq, 602, 003.
Discarding the terms which do not depend on 00s we have
1.
- 0 U)=2U •©• ct? +£ • 0- u;
Cc? M 2
Co\ + 2^1

s » ^2 cot $) C(J2 - OO1LO2)
.
(4.13.1
Multiplying the first expression by \M and adding the result to
eqs. (10) and (11) we obtain
5m«2 3 1
S* ~Co\ + 2^2 T 2^3 + 2002 001W3) +
(0O1OO3
2a; 3 (0^10^2 —
^1^2) — (cL^i aj2 ^1^2) cot $].

4.13.3 Two-axle trolley


While calculating the kinetic energy with discarded non-
holonomic constraints 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
ii,i2,ii,ii shown in Fig. 1.3, we have
VA = ii (-^1 sin x + OJ7 cos x) + i2^2, W4 = cos x, VB
= ii^7 + *2^1 > V5 = ii CJ7.
Let ui (and correspondingly Ui) denote the angular velocity of
the rear axle which is, due to Sec. 1.10, given by

ui = i3$ = y (cJicosx + ^7sinx-^2)i3, Ui = ycj7sinxi3, (4.13.13)

where i3 is the unit vector perpendicular to the road plane.


Vl = + Ui x i2 a = — iiai9,
Vi
a . \ y 1ii
= VA - sinX = W7 (cosx- sm
iiyW7 xj
V2 = v^4 Ui x i a = V4+ iia$,
2

v2 ' ii
= + iiyw7sinx = W7 (cosx + y sm
xj

respectively. The angular velocities having subscripts 3 and 4


are composed of the angular velocity of the axle Ui (and
correspondingly Ui) and the angular velocities of rotation
about the axle which are
Wi = £ , . a \ a [ sin x + 7 cos x)
wi + -UJ2 u>3 + (cosx - 7 sinx) W7 12,
l l
(4.13.14)

) OJi — — U>2 — ^4 + (
W2 - — - ^smx - y cosx cos x + ysinxj UJ 1
7 2,
(4.13.15)
provided that the constraints are discarded. When the
constraints are taken into account then u)\ = L02 — <^3 = <^4 = 0,
and we have
• Vi
u 3 = u i + 12^1 = h# + 12^ 1, U 3 = Ui H i2,
r 1
. . ... V2.
114 = Ui + 12^2 = ^3^ T *2(P2? U4 = Ui H-------------------i2,
r1

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

u2 = (tf + x) 13 = y (^1 cosX + ^7 sinX - ^2 + lus) 13,


U2 = y (W7smx + lv8)h,
where LU$ = The velocities of the wheel centres are
determined by the following equalities

v3 wB + U2 x i^c = V5 - c + xj ii,

V3 LJ7 (1 - y sinx) - cu;8 ii,

v4 wB - u2 x i'2c = vB + + xj ii,
LJ? 1 +
( 7 sinx
) +aj8
^
V4
T2<i>z -yWlCOSX+ yW2 ) U - CU>8, (4.13.1
-w5 +
CC 7

=- 6)

y <^i cos x - y w2 - w6 (4.13.1


^4 = +( ;i+£ShiX) UJ7 + cws- 7)

Now we proceed to construct an expression for the kinetic


energy. Let Mi and M2 denote the masses of the rear and the
front axles (without wheels) and 0i and ©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 mi and m2, respectively.
The tensors of inertia of the wheels at their centres are
respectively equal to
(iiii + i3i3) + Jii2i2, J3 (i'li'i + 1313) + J3i2i2.
J[
Now applying formulae (7.7) and (7.8) we obtain the following
expression for the kinetic energy of the rear axle

where — iis = ii AC. The kinetic energy of the wheels of the rear
axle is
+ T4 = —mi (^1 ^2) + 2+ 2^ (^1
Ts ^2) •
The kinetic energy of the front axle is
T2 = ~M2V% + —02 ($ + ^8^
whereas the kinetic energy of the wheels of the front axle is

+ ^6 = 2m3 (v3 + vl) + 2+ ^8) + 2^ +


^5 •
Collecting the above expressions we obtain the kinetic
energy of the trolley

2T (Mi + 2rai) (—u; 1 sinx + CJ7COSX)2+^2


+ 2MILU2$S +
(0i + 2mia 2 + 2 J’i) $ + ( M2 + 2m2) ( 'UJ2 T ^7) T (4.13.18) (02 +
2m^c + 2J3) ^
2
+ Ji {fp\ + <£2) + ^3 {fp\ + ^4) 5
4.13 Examples of kinetic energy and energy of accelerations 199
where d and Cps are given by eqs. (13)-(17).
If we take into account the non-holonomic constraints we
obtain the following expression
2T = (/i + sin2 x) ^7 + + 2U(JJ^UJ sinx, (4.13.19)

where
2 Ji 2J3
M + 2mi + A/2 H- 2 777.2 H 9—I- -o-,
rf r|
J2 (®i "I” 2TT7.I<22 + 2J{ + 02 + 27723c2 + 2J«Q
Mi
Mi - 2mi +

1 + 2?77.3C2 + 2^ +
v 7 02 (4.13.20)

As mentioned above, the constrains are taken into account


when calculating 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
dii TT • . <^7 sin x f = U, x i2 = (4.13.21)
— = u, X = ,2——,

di[ UJ 7 sin di'2 / u>7 sin x


dt = uix i' = a2 x + <^8 d7 = “ l l l — r~ +LV8
l (4.12.22)
We obtain the acceleration of point A and the angular
acceleration of the rear axle

wA = VA = ii (W7COSX - w7w8sinx) + ’12~jL sin x cos x,

Ui =i3y(w7sinx + W7W8cosx);
the acceleration of the wheel centres and the angular
accelerations of the wheels

Wi = Vi = ii ^77(c[ cos
UJ X - y sin x) - w7w(sin
8 X + y cos x)

. u>7 ( a.
12— ^cos x — y sin )i Wi
sinx = ii VFn +
2,
x 2
200 4. Basic dynamic quantities

w2 - v2 - ii w7 (cos X + J sin x) - 0J7V8 (sin X - y cos x) +


sin
h~^ (cosx + y sin x) X = 11W21 + hW 2 2 ,

u3 = Ui + — { h Wn - hw12), U4 = Ui + - (i W i - iiW ). 2 2 22

r1 r1
For the front axle
W B =V B = iicj7 - i'2 (y sinx +w7w8) ,

W3 = V3 = ^ ci;7 (l - y sinx) - y7u;8 cos +


x — CU>8
i2 (y sin x + ^s) [^7 (l - y sin x) - CUJ8 W3i/i'i + Wmi*

W4 = V4 = i'l W7 (l + y sin x) + -^7^8 cos x +■ CLJ 8


<
i2 (^j- sinx + o;7 ^1 + y sin x^ + cu8 = W4viy + W42/ i'2,

U2 = i3y (<Wsinx + <^ o;8 cosx + ^8,)


7

U5 = U2 + (l2^31' — I1W32') , Ug = U2 + yrj (*2^41' — ^\W42') •


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
wA - (£ Ui x sii — U\ ) 5
?sii +
S{ = -M\WA + Mi 0! Ui

Only terms containing ch7 and u>8 should be taken into account.
Thus, the second term in the latter equation drops out and we
have

S* = + 1 ( - Mi J (<h7 sin2 x + 2ta7ta7a;8 sin x cos x) •


Calculation for the front axle yields
1 10
*5*2 = -Mich7 + --y (th7 sin x + 2(h7a;7a;8 sinx cosx+ ^2^1+
2

21LO8W7W8 COS x + 2!cb8ci;7 sin x) •


4.13 Examples of kinetic energy and energy of accelerations 201

We have for the rear wheels that

S3 + S4 = 2 | (2mi
(wj sin 2
+
~r^)
x + 2UJ7W7LOS sin x cos x) | ,
2J[ 2Ji
P - 277li -f~
1-
and for the front wheels that

9 . 2J3 w7 + 2 3 cf
sz+ss*=~{l
6 2J3
- + p ( 2™3 + (th sin x
2 2
2 m3 H- -Y J!
+2UJ7U>7U>8 sin x cos 2 x + + 2lojgaj’—
r
Tuj8 cos x + 2Zci;8w7 sin

x) j • P

Adding these expressions and applying notation (20) we


obtain
2S* = (/x -b Hi sin2 x) ^7 + lvCd\ + 2VCJ7CJ% sin x c°s X +
2 II 1 L )7LU7UJ% sin x cos x + 2 ISUJ $ UJ 7L}$ cosx- (4.13.23)

The first three terms can be obtained by replacing quasi-


velocities a;7 and cos in the expression for the kinetic energy
with quasi-acceleration C07 and cjg, respectively. All other
terms are given in eq. (10.12).
These additional terms can be split into two groups. The
first group denoted by R\ 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 q.4 = x> by virtue of eq. (1.10.14)
only 643 = 0, while the other elements of the fourth row of
matrix b are identically equal to zero. Formula (10.9) takes the
form

k; = ihkA sr + hsA kr
[s
’ * * ~ hrA*sk) 5

ro
[8 M
, i 1 dAlr [s 8-rl = - ®Asr
' ' = YW' 1
’ ’ J*- 2 dX
r , QI IdKk
dA*8r
[8,8; r ], = dX ’ ► (4.13.24)
2 dX ’
r , r n r , 1 8 Ate
[s, 8; 8] , = [ 8, s ; 8] , = 0, [ 8, 8; 8], =

Now we can write the first group of terms


Ri= d;7(W2[7,7;7]T + 2w7a;8[7,8;7]x+a;2[8,8;7]J +
+w8 (w2 [7,7; 8], + 2W7U;8 [7,8; 8]w + UJ28 [8,8; 8]w) .
202 4. Basic dynamic quantities

By virtue of eq. (19)


A77 = M + Ml sin2 X, A78 = v sin x, ^88 =
and eq. (24) we can express Ri as
Ri = (2/i1cj7u;8 sin x cos X + VUJ\ cos
x) —
/ii^8<^7 sin x cos
X-
(4.13.25)
We proceed now to calculating the second group of
additional terms in eq. (10.12)

R2 = s^k^-lnir ~ U7CV8 £ £ A%lll%7uh +


rskl kl
LU8UJ7 ££ AklWkl\& ~ (^8^7 ~ tW<^8)
kl k

Here equalities (10.12) were used. The values of and are


obtained from eq. (18) for the kinetic energy
sin
^71 = Mi X cos x, -4.81 = v cos x-

Then we have
R2 = — LO7UJ8) (Mi^7 sin X c°s X + VUJ
8 cos x)
(4.13.26)
and finally
Ri + R2 = HIL07<JL)7<JL)8 sin x cos x + vw8<^7<^8 cos x-
(4.13.27)
Thus expression (23) is obtained. This calculation was
performed in order to demonstrate that it is possible to derive
an expression for the energy of accelerations without
knowledge of the acceleration of the system’s particles. To
achieve this it is sufficient to possess expressions for the
5

Work and potential energy

5.1 Generalised forces

The sum of the elementary work due to the application of


forces F* at points Mi of a system undergoing virtual
displacements <5ri of these points from their positions at a
fixed time instant is given by
N

6'W = J2Fi'Sri-(5.1.1)
2=1

The notation 6f 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 Sqs we obtain
N no n N p.

gw = I> ■'££■*. = X>. X> ■ £■■ (5.1.2)


2=1 S=1^ S = 1 2=1 ^
The quantity
N
Bv
«* = g F ‘ ' a i <5X3>
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
A. I. Lurie, Analytical Mechanics © Springer-
Verlag Berlin Heidelberg 2002
204 5. Work and potential

energy
of the generalised forces is given by
n
6'W = ^2Qs6qs. (5.1.4)
S= 1

Hence, the generalised force Q is equal to the coefficient of


s

variation 6q of the corresponding generalised coordinate in the


s

expression for the elementary 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 (6fW)k,
(<S'W)k=QkSqk,
(5.1.5)
and the generalised force Qk is determined as the factor of 6qk
in the latter expression. It is good practice to calculate all Qk in
this way.
When using quasi-coordinates we determine the virtual
n
N A n N ^ n

m = £ Fi • £ £-6* = E ^ E *' ^ = E P* ”" (5-L6)


s
F 6

2=1 S=1 S=1 2=1 S=1


The quantity
N s
^ 1 dns
(5.1.7)
2=1
is termed the generalised force corresponding to the quasi-coordinate ir s. Using
eq. (1.5.17) we can rearrange the latter equation, to obtain
Ps = = E 6- ' £ = il*rsQr. (5.1.8)
2=1 r=l r= 1 2=1
dqr r= 1
The inverse relationships are as follows
n
Qs = J2arsPr, (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 determined in general by the equation
r
i =ro+r' (qi,... ,qn;t) (5.1.10)
where TQ is a prescribed function of time. The virtual
displacement is, by definition, < 5r^, 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 Fi, F2..., FN- The virtual displacement of point Mi in the
body is given by the relationship

Sri = Sr0 + 6 x r', (5.2.1)


in which <5ro 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' = OM>i the position
vector of the point in question. Therefore,
N N N
S’W = ^Fi-^r^^Fi-tfro + ^Fi-^xr;)
2=1 2=1
N N
(5.2.2)

Since
N
(5.2.3)

is the resultant force of the system of forces under consideration and


N
(5.2.4)
206 5. Work and potential energy

is the principal moment about the pole O, we obtain the


expression for the elementary work in the form
S'W = V- Sr0 + m° • 0.

(5.2.5)
If we take the coordinates xo,yo, zo of the pole O with respect
to the fixed axes Oxyz and the Euler’s angles 'ip, $, (p as the
<5r0 = iiSx0 + i2%) + 13^0, 0 = i3Sip + H- i'3S(p.
(5.2.6)
Here is and i's denote the unit vectors of the fixed and moving
axes, respec-
tively, and n stands for the unit vector of the nodal axis.
Relationship (5)
reduces then to the form
S'W = ViSxo + VzSyo + V36Z0 + m ° • hS'ip + m ° • n<5$ + m° • i3&p.

(5.2.
7)
Due to eq. (1.4) we have
Qi = VL, Q2 = V2, QS = 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 = m° • i3 = ra3, Q5 = m° • n = mN, Q6 = m° • i'3 = ra3. (5.2.9)


Thus, the generalised forces corresponding to the Euler’s
angles *0, <p are
the projections of the principal moment on axis Oz, the nodal
axis and axis
Ozf which are the principal moments about these axes.
We now express the scalar products in eq. (5) in terms of
the projections (5.2.11)
of the vectors on axes Ox'y'z' fixed in the body
we obtain
Pi = mi, P2 : m2, P3 = m3, P4 = V1, 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 m\ instead of ra^, since the projection of
the moment m° 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 generalised forces are the
principal moments m i , m 2, m 3 about these axes.
Let us relate these quantities with the moments ra 3,rajv,ra3
which are the generalised forces corresponding to Euler’s
angles. We have
but, of course,
7^ m3i3 + mNn + ra3i3,
since the directions i3, n, i3 do not form an orthogonal trihedron.
The correct representation is as follows
m° = mNn-\——[(ra3 - ra3 cos d) i3 + (ra3 - ra3 cos d) i3l,

(5.2.14)
sin d
as can be proved easily by means of eq. (9). Indeed, scalar
multiplication of
eq. (14) by n, i3, 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 . sm ijjprojections on the axes is given in
m\ — mN cos yj +itsm3 ——-
Sec. B.2._ sm d
. . cos ^
= sm
Using eqs. (2.3.1)-(2.3.3) we m - - --
sin obtain, from eq. (14),
m2 rriN ip — 3
d
m3 = m 3
— rhs cot d sin
,~ ., , >
(5.2.15)
mN = idi cosip + rri2 sin^,
= sin d sin ^ — rfi2 sin d cos ^ + m3
1713 (5.2.16)
cos d, m3 = m3.
By analogy we have
„ sinw
mi = mNcos tp + m3---------T m3 cot ^sin(/;,
-
sin d (5.2.17)
~ . _ cos — m3 cot d
cp
m2 = —ITIN sm p + m3----- cos </?,
sin d
m3 = m3
208 5. Work and potential energy

and the inverse relationships


777,TV = 7711 COS (f — 7712 sin (f,
7773 = 7773, (5.2.18)
7713 = (777i Sin if + 7772 COS if) sin $ + 7773 COS 7?.

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 elementary work done by the
actual displacements as
N
d'w = J 2F
i'dri-

(5.3.1.)
2=1
Using this quantity we define the work of the forces during a
finite displacement of each point in the system as the following
integral
(2) (2) N
W 12 = jd'W = J^Fi- dn, (5.3.2)
(i) (i) i= 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 t\ and £2, 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
such that the projections of the force acting on the point
T/T, zf) are
equal to the negative partial derivatives of II with respect to the
corresponding coordinate, an p.i.e. _® an
=
. — ~
TP
dx
lV
dyP Fix = dzP (5.3.4)
IX P
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.
(2 ) (2 )
p d'W = - <m = iii - n2. (5.3.7)
J J
(i) (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
n = n(gi,... ,g n ). (5.3.8)
According to eq. (6) the elementary work of the potential
forces due to a virtual displacement is
22 r\TT
S'w = -sn = V
(5.3.9)
s=l °Qs
Comparing this with eq. (1.4) and taking into account that
the variations of the generalised coordinates are independent,
we arrive at the following
210 5. Work and potential energy

Assume that the generalised forces Qi, ...,Qn are known and
that they depend only on the generalised coordinates qi,...,qn. As
follows from the latter equation, an indication of a potential force is that
it meets the following condition
dQs dQk
dqk dqs (M 1, . . . , 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
Fx = —k J Fy — k
y
_____ x2 + y ,2 •
x2 + y2
It is easy to prove that condition (11) is met and that the
elementary work is given by
d'W= k (------ —-dx H——Aody) = fcdarctan —.
\x +y x +y )
z z A
x
Since function arctan(y/x) is not defined at the origin O, force F
is not potential in the region which includes point O.
Introducing quasi-coordinates we have, due to eq. (1.8),
n n
Ps b™Qr = - 53 brs 75-
rjyy (5.3.12)
oqr
r=1 r=1
or recalling notation (1.5.17)
PS = -|H (s = 1, . . . ,n).

(5.3.13)
U'K s
Sometimes the generalised forces can be represented by a
formal equality
(s = 1, . . . ,n), uQs
Qs = -TT!
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
n = n(qu ... ,qn,t).

(5.3.14)
This is particularly the case for nonstationary constraints
Fx = f( t )
5.3 Potential energy 211

Then
n ( M ) = -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
^ n OTT /OTT
d'w = Qsdqs = Y1 ird(l° = ~dn + ~^dt (5-3.15)
8=1 8 = 1 0QS 01

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 generalised 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 constant direction. For example, the gravitational
force in the neighbourhood of the earth’s surface. The
potential energy of the system in the gravitational field is
given by the expression
N

n = g ^2 m z
ii = Mgzc,

(5.3.16)
1=1
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 =
(5.3.17)
i=1

where g is the vector of free fall acceleration directed along


N

n = -J]FiTi.

As an example let us consider a system of two rods O 1 O2


and O2 O3
with joints at points 0\ and O2 , the free end O3 of the second
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 <p and <p between the rods and the axis 0\X. The angle
1 2

between the direction of force T and axis 0\x is denoted by (3.


The elementary work of force T is
8 W = TX6XO3 + Ty6yo3
f

=T [cos (38 (h cos (p 1 + h cos p2) + sin (38 (h sin cp 1 + l2 sin (p )\


2

and therefore
8'W = T [li sin (/3 — Vi) 8px + sin (/? — cp2)
8(p2], i.e. the generalised forces are equal to
Qi — Tli sin ((3 - , Q2 = Tl2 sin {(3 - ip2). (5.3.19)

These will be the generalised forces under condition (11), i.e.

h cos (13 - ip1)-^- = k cos (/3(5.3.


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 [li cos (/? - + l2 cos (/3 — cp )\. (5.3.21) 2

In the case of a force T having a constant value and a


constant angle a between its direction and the rod 0 0z, we have 2

(3 = a + ip and then
2

Q1 = Tli sin (a + <£2 — Pi), Q2 = Tl2 sin a.


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 O, namely an initial position
and a final position which is obtained from the initial one by a
finite rotation 6. The forces F ? , F ° acting on the body points
with the position vectors r5, . . . , 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
becomes, after rotation
+ 2^ X
Tfc = n + Ox
i + \e2

and by eq. (17)

k=1
n= Tfc (5.3.22)

(F°) +1# • F'


= n n- yk-e--e2 K
i + \e 2
L
k=1

The term IIQ , 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 O is given by
m° (F°) = 4 x F°.
Thus, the result can be cast in the form

(5 323)
n =
- TTW m S + ^T7We ^ '
Here denotes the principal moment of the system of forces
about the pole O in the initial position and Q° denotes the
following tensor

Q° = E ( E F 2 T 2 - F M ) . (5-3.24)
k—1

Expressing the vector 6 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

n = -2A0^Asm? *s + 2 (5.3.25)
S =1
Qsk^s^k-
214 5. Work and potential energy

Here stands for the unit vectors of the trihedron bound to the
body in
the initial position and Qsiz denotes the components of tensor
withfor example,
Q°Let,

Then,

and inserting the Rodrigues parameters (3.6.6) in terms of


Euler’s angles into eq. (25) we obtain
n= —2 (A0A2 + A1A3) FZQ + 2 (A2 + A3) Fx0
= —Fzo sin d sin 'ip + Fx0 (1 — cos 'ip cos p + sin 'ip sin p
cos i 9 ).
Of course, the same result can be obtained by means of the
n = —F (x — x0),

where XQ 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 = XQ (cos ip cos 11> — sin ip sin ip cos i)) + ZQ sin x) sin ip.

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 Ox
1X2X3 are prescribed by linear functions of the coordinates

Fs = PgiXi + Ps2X2 +Ps3X3 (s = 1 , 2, 3 ) , (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.4.10). Let P' denotes the transpose of tensor P, i.e. such
a tensor that Pik — Pki• The relationship
o —
P') = l Qs
0 -a
{ -n2
0

$^
3
^2 =^ (Pl3~ Ps l), =

i> (5.4.6)

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
-f]3X2 + ^2X3, f]3X i ~ f ] i X 3 , -fi2Zl + ^1^2-
We introduce into consideration a vector Q, whose
projections on axes OX1X2X3 are respectively then

F = S-r + ftxr.
(5.4.8)
This notation is justified by the fact that under the
transformation of the coordinate system values 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
1 i3 3
-II=-r-S-r=-^^ S i k XiX k ,

(5.4.9)
i=1 fc= 1

i.e. it is the potential part of the force F


216 5. Work and potential energy

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 vector of reaction of the oil film on the rotating shaft
is an example of a circulatory force. By Sommerfeld’s theory
127T/IR3
F = zu> x r. (5.4.12)
e2 (2 + A2) \/l —A:
Here Rf and R denote the radius of the bearing and the shaft,
respectively, e = Rf - i?, fi is a factor of the lubrication viscosity,
A = e/r, and UJ 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 O'.
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° , the latter being proportional,
and in the opposite direction to, the vector u. Thus the point at
which the resultant of the oil film reaction forces is applied

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 (5.5.1)
_ mmor
/ o 5
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 / is equal to
/ = 6.67 • 10 11 nr? I kg • s2. (5.5.2)
The elementary work of the force F due to the virtual displacement 6r
is
c/rrr pTflTTlQ
6 W = -f—r • 6r.

Since
r • <5r =-6 (r • r) = -6rv2 =
r<5r, 2 2 ’

we have
6 'W = -f^6r = 6(f^).

It follows from eq. (3.9) that the potential energy of the force of attraction of
two particles is described by the expression
rara (5.5.3)
n = -/ o
r
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 ofn eq. m(3)
= -/ E— - r (5-5-4)
i

where ra* denotes the mass of particle Mi in the attracting


body, and r* the distance from the attracted particle M. The
position of the particle Mi in the body is given by the position
vector p{ = GM>i, see Fig. 5.3. Denoting r = GM we obtain
= r - r\ = r2 + p? - 2r • pi
and
1_
Ti (5.5.5)
Furthermore

r ■ p i= x xi + y yi +z zi , p f = x f + y f + z f ,
218 5. Work and potential energy

FIGURE 5.3.
where x , y , z and x^y^Zi denote the coordinates of points M and
respectively. Expanding (5) as a second order polynomial in
p/r, this ratio
being assumed to be small, we arrive at the following
equality
2 =
1
(■, , T_Pi 1 Pi | 3 ( r ~ P »)2
n y r2 2 r 2 r
2 4
r

_ 1 xxj + yyi + zzj 1^ + yf + z? 3 ( x x j + yyi + z z j )2


r r3 2 r5 2
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
^rriiXi = ^ra^ = 0,
i=1 z=l z=l
whereas the coordinate axes coincide with the principal axes
of inertia such
that
N ? - oL 13 N N2 + *2) + |^B 13
+ y*
n -fm
i=l i=1
N
3 y^ v—^ o 3z2 ^^ o

z=l z—1
Here mo denotes the total mass of the body. Due to eq.
(4.3.21)

m<
(Xi + Vi +
Z— 1
53 ** ) = ^ (©1 + ©2 + ©3) ,
an 'an fm 'X '
(02 + 03- -20i)
dx dr 7*4 r’
an 'an fm (0 + 0i- V
3 -202)
dy dr 7*4 r
an 'an fm
(0! + 02- -203)
z
i
dz dr 7*4 r

where
dUfm + 2^6 K®2 + ^3 - 2©i) 'r" + (03 + 0i - 202) y2+
&=
(01 + 02 — 203) 22] } . (5.5.9)
The principal moments about these axes are
given by
m,r ■ yFz - zFy 3 fm'■ (03 - ©2) yz,
z F x - x F z = ^ (e 1- e 3) z x ,
(5.5.10)
Sfm (n n x
— — (02 - 0i) x y .
m. X
Fy yFx

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

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 M(5 = —r or by the
coordinates £, 77, £. The directions of axes Gxyz are described
by Euler’s angles which determine the direction cosines of
these axes with respect to the fixed axes.
Projecting vector r onto axes Gxyz we have
(5.5.11)

These relationships should be substituted into expression ( 6)


for the potential 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 ellipsoid 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 6 between axis Gz and vector GA$ = r is positive, see
Fig. 5.4b.

Due to eq. (7) the vector m G , which is perpendicular to r , is


now perpendicular to axis z, i.e. the unit vector i'3. This can be
cast in the form
m (5.5.13)
\
Fx = fm ^ +( 0 3 - 0 1 ) ( 1 - 5 c o s 2 9
)
X
5
r
5 +JI(03-ei)(l-5cos20) y
-, >
II

^ + ^ 4 (03 - 0l ) (3 - ° °)
5C S2
c o s 9.
Fz = fm
j
>
0 11

sin#,
£
II

^ + 1^(1-500^) _>
II

COS0,
^ + !^74^(3-5 COS 2 0)' >

We now enter the orthogonal trihedron of the unit vectors n, n',^


”half- fixed” in the body. The vector n is orthogonal to the plane
of r and i(j, while n' is directed along the component of r in the
plane Gxy as shown in Fig. 5.4b. Then

(5.5.15)

3 fm
mn (03 - 0i) sin 2(9, mnf = 0, mz = 0. (5.5.16)
r 2 3

In the case depicted in Fig. 5.4b, i.e. ©3 < 0i, the moment mn < 0.

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 [40].
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, $, A of a point denote the
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 = ©i = ©2
and C = ©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
1 C-A r 2Mr5
III -fM (x2 + y2
1 C-A r 2Mr3
-fM (sin2 & 2 cos2 #) (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
n2 = — ( x2 + y )
2
= — sin2 #,
(5.6.2)

with U being the angular velocity of the earth. Summing up


eqs. (1) and
/ M k C-A (ROY (Z_\
n Ro I r + 2MRl V r ) + 2fM \ R o )
3 {C-A) Ro\3 U2Rl r )
+
2 MRl 2fM

The following notation

U2Rl 3 ( C - A ) U 2R Q
(5.6.3)
fM ~ m’ 2MRl + 2f M

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
5.6 The shape of the Earth 223

This is equivalent to the expression

(5.6.5)

in which P2 (cos ti) denotes the second polynomial of Legendre


P2 (cos'd) = - (3cos2$ — l) .
(5.6.6)
In what follows, the products and the squares 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.
fM
(5.6.7)
~W~
90

represents the acceleration of an ” average gravity force”


which is the attraction force of the non-rotating earth at radius
Ro. With this in view the value
U2Ro
do (5.6.8)

is the ratio of the centripetal force on equator to the average


gravity force.
The above value Ro is now defined as the radius Eo of such
a sphere that its specific potential energy IIo is equal to the
average value of the potential energy on the surface of this
1
n0 4:7TR nr=ZjR0do,
Q
II

where
do = RQ sin tidtidX
denotes the element of the sphere surface EQ. Hence,
due to eq. (5)
n0 = j d\ IR20 nr=jRo sin Md
4TT Rl
27T 7T

fM
fM
2R0 /(l sin'tfdf? — —a
+ J P2 (costi) si sin tid'd Ro )
224 5. Work and potential energy

since by virtue of the Legendre polynomials


7r

J P2 (cos T?) sin tid'd = 0.


0
Instead of II we will consider its deviation from IIQ

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 Ho which is an ” average spherical
earth”. This definition introduces the distance R* between the
centre and the surface of the normal spheroid as follows
(n - n0)rz=jR* = 0.

(5.6.10)
R* = Ro [1 +7 ( 0) ] -
The quantity 7 (ti) 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),
—7 (ti)
— -a+2 ( ?) = 0, COST

o
and the equation for the surface of the normal spheroid takes
R* = Ro 1 — -a+2 (5.6.11)
(COST?) o

At the equator and the poles, i.e. at i9 = 7r/2 and i9 = 0

(R*)^/2 = a = R o ( l + j ) , (R*)d=o = c = RQ (l — |<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
RQ 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 P 2O?)
= 0, i.e. at
COST? = ±^1, i?i = 35°16', i?2 = 144° 44'.
The equation for the normal spheroid can also be written in the
form
R* = Ro ^1 + ^ — acos2d^ = a (l — acos2d) .

(5.6.13)
Let d* 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 coordinates of the point, then on the
dz* 1 — 3a cos2 d
tan d = —-— = tan d « (1 — 2a) tan d
dp* 1 + 2a 3a cos2 d
or
d* - d
tand* — tand «--------— = —2a tand, e — d — d* = a sin 2d.
--------------------------(5.6.14)
cos^ d
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
m RQ
2 r
Fr = - dr ~9o r — -m——
3 Ro
R4 2r
(2a — m) —jP2 (cosd) + - m — P 2 (cosd)
V (j iio (5.6.15)
p dU
L2d,
4
f = / m\ (Ro\ m r
» -7M = 9° +
V T H T J T jfc
PA = 0,
where go is given by eq. (7). At the spheroid’s points we have
. a3 ( 5 \ 9 .
P* ~ ~9o 1 + — cos d
— —rri - la - -m )
(5.6.16)
F£ = g0a sin 2d.
The components of the gravity force along the normal n to
the spheroid and the tangent r to the meridian are as follows
F* = F* cos s — sin £ & F*,
F* = Pr* sin e + F# cos e — —goOL sin 2d + P^ = 0.
226 5. Work and potential energy

The latter result serves to validate the calculation validity


because the vector gradll is normal to the surfaces II = const
and the spheroid is bounded by such a surface. The
expression for F* can also be written in the form
K = -ge (l + /3cos2tf) ,
(5.6.17)
where

< fe = S d ( l + f-§ro)

(5.6.18)
stands for acceleration of the gravity force at the equator and
the Clairaut constant is
(3 = — a.
(5.6.19)
We also construct an equation for the normal spheroid by
using cylindrical coordinates. We have due to eq. (13)
R*2 = p*2 + z*2 = a2 (l - Oi cos2 $)2 « a2 (l — 2a cos2 #) « a2 - 2az*
or
By virtue of eq. (12)
a2 « c2 (1 + 2a)
and thus
+ 4- = 1.

(5.6.20)
a 2 c2
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
a = 6378.4 km, a » ~ 0.003375.
(5.6.21)
5.7 Elastic forces 227

By means of (3) we obtain


A
3(7-
(5.6.23)
0.00164.
2 MRl
The equinox anticipation theory yields
C—A
——— = 0.003275. (5.6.24)
O
Comparing the latter two equations we obtain
C
MR 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 7T/2 — and the
geocentric 7T/2 — d longitudes is given by eq. (14). Its maximum
values is achieved at longitude 45° and is equal to a. A more
accurate analysis yields £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
Qs — Qs (shi • • • 5 Qn) •
(5.7.1)
are potential forces. This physical assumption
Elastic forces
expresses the property of elastic bodies to accumulate
potential energy under a monotonic loading and to give it back
without any loss when the body, under a monotonic
Cll Cl2 • • • Cln
C21 C22 • •

Cnl Cn2 • • ■ • Cnn

c= (5.7.6)
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 (fc,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
n n
OTT n n
fll = 53 f^Sq, = - 53 QaSq, = ££ CskQk^Qs-
(5.7.8)
s=1 s—1 k=1 s=1

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)

11
= \ £ fp?s = 53 Qsqs = -\0>q = ~\q'Q, (5.7.9)
s=l S=1

then, by virtue of eq. (5), we obtain


-j n n
n
= 2 £ £ Cak a
^ =
2 ' '
q cq

(5.7.10)
s=1 k=1 Z

Here q and q are a column matrix and row matrix of the


f

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
an Oi\2 • • • Oil n
c-^
&21 &22 • • • Ot2n

Qnl OLn2 • • • OLnn

(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
=
^ ^ C^skQk — (&slQl H • • • &snQn) ($
-
1, . . • , Ti) .
(5.7.12)
k=1
The inverse of a symmetric matrix is also symmetric, i.e.
Olsk — &ks $ — 1 , . . . , Tl) . (5.7.13)
This equation expresses Maxwell’sreciprocal theorem.
Substituting qs in terms of the generalised forces into a
bilinear expression for the potential energy (9) yields another
representation for II in terms of the generalised forces
.. n .. n n
n = - 21 X) = 2 E E 1a'kQsQk = -jQ'aQ- (5.7.14)
S=1 S = 1 fc=1
By virtue of eqs. (12) and (14) we arrive at Castigliano’s theorem
an
n
^ ^ OiskQk — Qs (5 = 1, . . . , 7l) (5.7.15)
dQ . k=1
s

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 elaborated in this and the
following sections.

5.8.1 A statically determinate system


A rigid plate S attached to two trusses at points Oi and O2 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 xGi yG of the pole G are taken as the
232 5. Work and potential energy

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
XQ = xGCOS <P + yGsin (p&XG+ Da = ~XG sin <p
+ VGcos « -xGy + yc,
differ from XQ 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\ and O2
on axes G^xy are equal to
vox — %G + bp, yox =yG- cup, xo2 = xG - bp, yo2 = VG + 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 xox, yox and xo2,Vo2, 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
denote the force in the k — th rod. By analogy we find the
forces /ik due to the unit load Y — 1 shown in Fig. 5.8. Given Xk
and fik, we find the forces Sk in the rods of the truss subject to
arbitrary forces X and Y at hinge O

FIGURE 5.7.
5.8 Calculation of the potential energy for rod structures
233

FIGURE 5.8.
as well as the potential energy of the truss as the sum of the
potential energies of all m rods
i 171 c2 i/ 171
»2 m> m2N
wr+2xyEwF-+ylEwir
k=1 k=1
(5.8.3)
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
\2 = E EkF^ A
(5.8.4)
E -E
an k=1 A .
EkFk ’ C*12 ^22 EkF
L
k=1
k k=1

Since the external forces Y and F can be understood as the


generalised forces —Qi and — Q2 corresponding to the
generalised coordinates Xo and yo (horizontal and vertical
displacements of the hinge O), we obtain
n — — (anQi + 2(X12Q1Q2 + ^22^2) • (5.8.5)

Because
dU dU
X0 — (anQi + CK12Q2), yo = dQ2 ■ (0>2lQl + &22Q2)
dQi
the expression for n in terms of the generalised coordinates XQ
and yo takes the form
n = \(cnxl + 2cux0yo + c22yl),
(5.8.6)

where the coefficients of the stiffness matrix are


C
1 1 1
11 T—T 22,Q
CI = —T—rO'12, C 2 = -j—fCU m
a 2
a a 2
234 5. Work and potential energy

FIGURE 5.9.
Returning to the example of Fig. 5.6 we have

where c'ik and c"k are coefficients of the stiffness matrices of


the left and right trusses, respectively. Expressing by means
of eq. (1) #Oi, •••, Z/o2 m terms of the generalised coordinates q\
— XQ, Q2 = 2/G, #3 = V we obtain
(5.8.8)
s=l k=1
where
C\\ — c'n + C J J , C22 — C22 ~b C22, \
C33 - b2Cu + a C22 + 2ab (c'12 - c"2),
2
(5.8.9)
C12 — d\2 + ci2i C23 = —bC12 + a (C22 - C22)
C31 = -bCn + a (ci'2 - c'12).

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 Z?,..., Z° and the direction is given by
the cosines ak,fik,^k of the angles
between this direction and the axes M$xyz fixed in space. The
potential energy of the elastic forces under small
displacement u = MQM of point MQ to point M with the
coordinates x, y, z is sought.
Under the above displacement the rod rotates through a
5.8 Calculation of the potential energy for rod structures 235

and neglecting second order terms we can adopt that the


vertex angle Mf of the triangle MpM'M is the right angle. Then,
denoting the unit vector of direction OkMo as e° we have
fk = U • e°k = akx + (3ky + 7 kz
and the potential energy becomes
^n 1 n
n = - ^2 Ck ti = 2 XICk ( akX + P ky + lkZ ?
k=1 k=1
where the axial rigidity of the k — th rod is given by
EkFk
Ck — i°k

Therefore
n= 1 {c\iX 2 + C22V2 + C33z2 + 2ci2xy + 2c23yz + 2c31zx)

\ \\xyz\\c (5.8.10)

where
Cll = c
12 = E Ckakpk C13 = E k&klk
k®l
c c

k=1 k= 1 k=l
n
C22 = £ ck0k (5.8.11)
C21 = ^23 E
k=1 C32
C12 Ckfik'Yk
= C23 k=1
n
C33 = E ckil
Let us consider now the same system of rods, but under
the assumption that the rod lengths lk due to assembly differ
from the initial lengths lk. Denoting as above the elongation of
the k — th rod by fk we have
Lk — Ik = (Lk — h) + (h — Ik) — fk +

where 6k stands for the change in length caused by system


assembly, with positive Sk corresponding to the tension. The
potential energy of the k — th rod is equal to
\ck (Lk - Ilf = l-ck {ft + 2 fk6k + St) .

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
Cn Cl 2 C 13
c =
C 21 C22 C23
C 3I C 32 C 33

(5.8.14)
5.8 Calculation of the potential energy for rod structures 237

|A/vwyflyww'cj|
X
- 21--------------------------

FIGURE
5.10.
with the entries given by

^11 =
X]Ck + +
7fc) fc=l
$k_
C12 S^H)’ Ik

Cl3 S^H)'
E 01+^ (71+ a i )
^22
C23 /c=l
X>/?fc7fc(i / * ) >

C33
7fc + -^ («fc + /3fc)
^21 Ecfc
k= 1
— Cl2, ^32 — C23, ^13 — ^31-
Notice that the quantity
Fk = ck6k

(5.8.15)
is equal to the initial stress in the k - th rod. Thus R1,R2,R3
represent the projections of the resultant vector of forces at
the joint Mo. Provided that there is no external force applied
Ri = 0, R2 = 0, Rs = 0.
(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.
238 5. Work and potential energy

FIGURE 5.11.
compression of each spring by 6 we have
ai = 1, a2 = -1, 0, 61=62 = -$,
h= h = hci = c2 = c

and the potential energy takes the form

n=c (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 O 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^r]( are fixed to. We look for the
potential energy of the elastic reaction of the rod under the
rotation 0 and displacement u of the body S. The final position
of the trihedron M£rj£ is designated by 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 a,/?, 7 of the
rotation vector 0on axes M^rjQ will be taken as small quantities
so that their products will be neglected. The rotation about
axes M£, M77, M( through the angles a, /?, 7 makes the
trihedron parallel to the trihedron The order of
u = qi, v = q2, w — g3, a = q4l P = qs, 7 = 96 (5.9.1)
5.9 The potential energy of a rod under bending, torsion and compression
239

FIGURE 5.12.
are adopted as the generalised coordinates. The
corresponding generalised reactions are the projections of
the resultant vector V° and the resultant moment L° of the
elastic reaction in the rod

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 of the rod is
in
equilibrium under force — V° and moment — L° in cross-
section M and V and Lin cross-section A-. The equilibrium
V - V° = 0, L —L° + r ( s ) x V = 0
or

V = V ° , L = L ° — r ( s ) x V °. (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 are designated
by
£ (s), r] (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 position
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 crosssection, and axis Az along
the tangent to the rod axis. The direction cosines ctik of the
angles between the axes Axyz and axes M^r]( are prescribed,
too.
Projecting the second equation in (3) on axes M£rj( we
L5 = L ° - r ? ( s ) y ° + C ( s K , '
Lv = L°-C(s)Vi0 + ((s)V<0, (5.9.4)

>
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 — ii + L^Oiyi + |
Ly — 2i + Lrja22 + L^a 23, > (5.9.5)
Lz — L^a31 + £77^32 + L^a33. J
The potential energy of the rod is composed of the
potential energies of bending and torsion

Here A = EIX and B = EIy 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,Lz from eqs. (5) and (4) into (6) and
replacing ..., V^° by the generalised forces (2) we come to the
following equation for n as a quadratic form of the
generalised forces
1 66

"= EE P skQsQki

(5.9.7)
s= 1 k= 1
where the 21 elements (3sk of the influence matrix (3 are the
definite integrals of some functions of s depending upon the
A-1 0 0
a= 0 B~l 0
0 0 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'0a' + V&a',


we obtain
L'aL = L'0a'aaLo + VQ^a'aaLo — L'0a'aa^Vo — V^a’ aa\Vo. (5.9.11)
Introducing now the following matrices
i i i
= aW£*, ! = - / * # ,
(5.9.12)
000

we can rewrite eq. (6) for the potential energy in the form
i
\ J L'aLds = 1 (yifcVo + V^L0
n = +

L'oltheVsecond
where 0 + L'and 32L
0(the 5 9 13 are equal to each other.
third terms 0 ) , ( . . )

In order to derive representation (7) we introduce the 6x1


column matrix of the generalised forces
^o
L (5.9.14)
0i 7'
7 @2

£ V
rf
c
X
y
0
-rfand the0 following 6x6 symmetric
z 0
a
matrix
0 1
(5.9.15)
0=

Because

Q'0Q = 0i i7 Vo PiVo+ 7^0


nL'0 74
02 Lo 7^o + P2L0
= Vo0i Vo + VoVLo + L'olV0 + L'O02LO

we obtain by virtue of (13)


1
n = :Q'PQ, (5.9.16)

which is the desired relationship (7). Formulae (12) and (15)


show how to calculate the influence matrix /?.
We proceed now to some special cases.

5.9.1 Plain curve


The calculation is simplified when the curve axis in the natural
configuration 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 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£r/ and z parallel to axis £. The equations of
the curve are given by
£ = £(«), v = v(s), C=0

where a prime denotes differentiation with respect to 5.


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
LZ = L ( L < = L°C-Z(8)V2 + V(B)V?. (5.9.17)
f rj2ds f &ds J f rjds
MV = J Elz ’ /J<D
Pl2 “ — _ Elz ’ 0$ = J Elz
0 0 0 > (5.9.2
i „ if kis if ds 2)

M
'-'C
CO
^
022 = J Elz ’ 033 J Elz
=

II
1
J Elz ’

000
^(N reds
^<N

II J Elz
0
£
re#
J Ely 0
II

1 1

> (5.9.24)

The other elements are identically equal to zero. If we


assume that the rod is extensible it can be shown that

_ EF’
11 — /
P(i) (5.9.25)
5.9 The potential energy of a rod under bending, torsion and245
compression

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
Hi = \^Ql '
n2 = i (f^Ql + 20g]Q2Q3 + pWQl),
(5.9.26)
n3 = ~ (fflQl + 2f3^Q4Q5 + 0%Qi) ,
n4 =
Here U2 and n3 correspond to bending in the planes £77 and
£77, respectively, ni and n4 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 0£ be the cylinder
axis and axes 0£, O 77 be perpendicular to this axis, see Fig.
5.13. The equations of the axial line of the spring are as
follows
. s cos . s cos a
a 77 = a sm-------------- £ = s sin a, (5.9.27)
£ = a cos----- 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 0£T7 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° applied at
point O and a moment L°. The displacement of point O and the
rotation angle of the plate are sought under the assumption of
small displacements and rotations. The axis Az coincides with
a y c
X — cos ip — sin^? 0
y sin a sin ip — sin a cos cos a
z — cos a sin cp
cos a cos cp sum
ip

where (p = s cos a/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
V? =the
Qi,axial
L\ = Q .
2
moment
The corresponding generalised coordinates

Qi = w, q2=l
are the displacement of plate S along axis and the rotation
angle about this axis, respectively.
The calculation yields
—aV® sin ip, Lv = aV® cos <p, L^ = L°,
0, Ly = —aV® sin a + L® cos a, Lz = aV® cos a + L® sin a
and
0n = la2 sin2 a cos2 a
+
B C~
sin a cos (5.9.28)
012 = la [ 77 - o
c B a:,
022 cos + sin2 a
2

a ~~C
B ~
5.10 Power 247

The inverse matrix is


Cn = ~r~2 {p cos2 a
+ B sin2 a) ci2 = j- (B — C) si
sin a cos a
la2 la
c=
1
C21 = C2 2 = y (C sin2 a + B cos2 a)
(5.9.29)
C12
Therefore the generalised elastic reactions of the spring are equal to

C cos2 a + B sin2 a B — C
Q1 -w — - 7 sin a cos a,
la2 la (5.9.30)
n B~C C sin a + B cos2 a
2

Q2 =----------:-----w cos a sin a —
la l "7-

For a wire with circular cross-section


nEr4 7T Er4
B= C=
4(1 + 1 /)’

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 v sm a cos a w
— U W . IS OIll UL LUO LX UJ . o
-1 \
7 =-----------------------------------sin a cos a = ----------------------------. (5.9.31)
C sin a + B cos2 a a 1 + v cos^ a a
Similarly, applying only a torque results not only in rotation 7 but also in axial
(C — B) sin a cos — a v sm a cos a
W = n2 z , D • 2 o^- -^7
’1 - -2 aj. (5.9.32)
G cos a + B sm a 1 + v sm 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.

A, _ E F ( . V i = 2>(E|£i. + t
2=1 2=1 \S=1 1
)
(5.10.1)

or by definition (1.3) of the generalised forces


n N

dr A
N = ^QSQS + E ^^ f
(5.10.2)
s= 1 2=1
dwi 1d ldv? dvi
Vi •Vi = = Vi —
dqs ~ 2 dqs 2 dqs uQs

Noticing that
5.11 The dissipation function 249

we find that
TV
d d<
r,

Qs = - (Vi)
dqs
fi (u) du = t> (s= 1, . . . ,n),
*=l OQs
dqs
(5.11.2)
where $ denotes the dissipation function

$= fi (u) du. (5.11.3)

Clearly, 4> > 0 since all the integrands are positive.


The generalised forces corresponding to the quasi-
coordinates are, due to eq. (1.8), equal to
n
d<f dus
P§ — ^ ^ brsQr (s = 1, . . . , n ) ,
5> > (5.11.4)
r=1 r= 1
d(}
where LJ denote quasi-velocities.
S

The dissipation function was introduced by Rayleigh in his


classical treatise [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
f(Vi) = V? (5.11.5)
the dissipation function

TV TV n 0 m+1
$= UTi (5.11.6)
m ++ 1 1=1 1
. m + If-'1
1=1
E
-x-Qs
is a homogeneous function of degree (ra + 1) in the
generalised coordinates. Using Euler’s theorem on
homogeneous functions (4.1.12) we can easily relate the
dissipation function
n
to the npower of the dissipative forces
N = y2Qsqs = ~ y 2 — q s = (5.11.7)

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 necessary 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
n — 1 of the velocity and
2s+l

Q = -kq2s ^ = -kq2s sign q, dq

where sign q = +1 if q > 0 and sign q = — 1 if q < 0.


Let us consider the motion of the system relative to moving
axes Ox' y'z'. Here we assume that expressions for the position
vectors of the system points with respect to the pole O do not
contain time explicitly. The resisting forces acting on the system
points are determined by their velocities relative to the environment which do not
coincide with the velocities relative to the basis Ox'y'z'.
For example, let us consider motion of a mathematical
pendulum whose velocity of the attachment point is VQ 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
VQ + v' whereas its velocity relative to the flow is v 0 + v' — V.
The latter expression for the velocity must appear in the
equation for the resisting force
F = —kf (|v0 + v' — V|) vp + v' — V |
v0 + v' - V|
'
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 VQ =
v' = 0 the point is subject to the force
F = kf(V)y

directed along the air flow.


Returning to the general definition, we rewrite equation ( 1)
for the resisting force in the form

(5.11.8)

where v* denotes the geometric difference of the vectors of


translation velocity and absolute velocity of the environment
medium V* at that point of the moving system where point Mi
takes place instantaneously. Vectors v*
5.11 The dissipation function 251

can depend on time t and the generalised coordinates


determining the position 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, v* 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
dqs <9v'
=
® / * i\
(v +Vl)
% iS ' -

because v* is independent of the generalised velocities.


Repeating the derivation of formula (3) we obtain
N <+tA(v*+v')
Qs X^/d^ +Vil) /I V ^
dq s
2= \v*i+v'i\

1
N
X kif (K* + v i D lv* + vil = (5-11.9)
2— 1

where now
|t>* +v[ |
Nki" "
$ = ^2 / (u) du. (5.11.10)
2=1 i

Provided that dependence (5) holds, the dissipation


function remains a homogeneous function of qs only if v* = 0.
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 == Fi
2 1
X
=1
• V
* = X
22=1
F
* '
V
i + X F
* ' V
“’

where v* and v^e denote vectors of absolute and translation


velocities, respectively. The first term describes the virtual
power given by eq. (5.10). We obtain
N v ?+ v (
N ■ XI ®s<is ~ X kif (K + Vi\)k * + (5.11.11)
s =1 2=1 v'i\
where Qs is given by formula (9).
252 5. Work and potential energy

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 O moves with a velocity VQ 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 translationally with velocity Vo. The squares of the
absolute velocities of the points are given by
v
i — l i + o| = (v0 cos a - sin^)2 + (vo sum + cos^)2 = vl + l\Cp\ + 2v0li<p1 sin (a -
v v 2

,
(5.12.1)
v\ | 2 + VQ| = (VQ cos a — l\(px sin^ — 1^2 sin<^2) +
= v 2 2

(5.12.2)
= v\ + + 2v$l2<i>2 ' ( ~ ¥2) + 2ZIZ <^ ^2 (Vl “ ^2) •
S n a
2 1
COS

Since
u

0
the dissipation function, due to eq. (11.10), is as
follows
(5.12.3)
5.12 Examples of the calculation of the dissipation function 253

The generalised forces are equal to


Qi = -kiVi [ifip! +V li sin (a -
0 '
k2v2 + v0h sin (a - px) + hl2<p2 cos - p2)\ ►
Q2= -k2v2 [l2<p2 + Vorsin (a - <p2) + hhP\ cos(</71 - <p2)\ .
(5.12.
4)
In the relative equilibrium Cpx — (p2 = 0 and the generalised
forces take the form
Qi = (h + k2) VQII sin (<p$ - a) , Q2 = k2v^l2 sin - a) ,
which is easy to prove by direct calculation of the elementary
work of forces -kiVQVQ (i = 1, 2) due to virtual displacements <Sri
and Sr2. The elementary work of the gravity force in these
virtual displacements equals
sin</?5<Vi — P2 (^1 sin</?5<Vi +^2 sin (p2S(p2) ,
—P\h
and the generalised forces due to gravity are
Qi — ~ [Pi + P2) h sin (fi, Q2 = -P2I2 siny?2-
From the equilibrium equations (see Sec. 6.5)

Qi + Qi — 0, Q + Q 2 — 0,
2

we obtain the values of <p\ and p)\

cot (fi P1 + P 2
cot a — cot if2 = cot a sin a*
(ki + k2) VQ sin a’ k2 VQ
(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
E\ — (p1 — ip® and e2 — ip2 — ip2 as well as the angular velocities k\
and e2 to be small quantities. Then, neglecting the products
and the squares of these quantities and taking for simplicity a
= TT/2, we obtain
and furthermore
ki + k2 ,tan p>2
0 k2 2
= ~—v 0.
tan pi = —
P1+P2
In view of these equalities we have
ki + k2 2 COS p>2 11k
cos cos ifl
22
1+£1
prTft”” COS if 2 l+e2—v0
254 5. Work and potential energy

FIGURE 5.15.

cos(<Pi - ¥>2) = COStp°COS<p%(fci + hz )^2 4


. + (Pl +P 2 )P2 V °.

Inserting these equations into eq. (4) yields

Q1 (fci + fc2) v%h COS(/?y 1 + ki£1+


-5-fc;„,2 h
2

P1 + P 2 + — x
Vo
(fci + k 2) k2 4
(l + COS (fi) £1] — 2k COS tp® COS (/?2 1 + - , „
2 1
„ ( „ v?
2(P1+P2)P2 °_ ^2,
Q2 = —k2Volil2 COS (Pi COS 1 (fci + fe) k2
ip® 1 + -2 (P1 + P2)P2 °. £1-

k2vll2 cos ^( 2 ^ oJ+^( +


1+£ t, 1
COS' ¥>2) ^2
(5.12.6)
These formulae are meaningful only for non-zero and
sufficiently large VQ since if VQ = 0 the expressions for v\ and v\
become quadratic forms of t\ and £2-

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 LJ about the pole O
Z

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
—fpdo—,
V
where / is the coefficient of friction which is assumed to be
independent of velocity, and p is the pressure which is
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.

$= (5.12.7)

We have
v = vo + L J x r
or in projections on axes Oxy bound to the plate
Vx = Vox - uzy, Vy = Voy + U Z X.

Therefore
$ fp JJ \J ( - u z yf Vox + ( V0y +

wzx)2dc
s (5.12.8)

Here
JJ \j(xp - xf (yp - yf do.
fp\vz\ +

v
0y Vo
Xp =---------
yp x (5.12.9)
Wz ' UJZ
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
S = /G>*|r*, (5.12.10)
where r* has the dimension of length and is equal to

r* = JJ5 rdo. (5.12.11)

Taking the instantaneous centre of velocity as the origin of


the polar
coordinate system (r, A) with the polar axis OP we have do = r
dr dX.
Hence, if the instantaneous centre is inside of S then as Fig.
5.16a shows
27r r3
r* = ^ J (\)d\, (5.12.12)
0
where r (A) = PM, M being a generic point on the border. If it
is outside r (5.12.13)
S', then, as Fig. 5.16b displays,
A2
256 5. Work and potential energy

FIGURE 5.16.
where 7*2 (A) = PN, r\ (A) = PM and the angles Aiand A 2 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 = cos A + Va2 - R2 sin2 A^ = R3 cos3 A + (a2 - R2 sin2 A)3^2
+3i?2 cos2 A Va2 — R2 sin2 A + 3R cos A (a2 — R2 sin2 A),
where R = OP. As the integrals corresponding to the first and
the last terms equal zero, we obtain
_4 R2 R2 K I*
_ 7+ E 4 1- a (5.12.14)
97
where K and E are the complete elliptic integrals of the first
and second kind with the modulus
1 < A; = — < 1.
a
In the second case, see Fig. 5.17b, we have

r2 - n
-rf (r2
= -ri) 4 ri + rir 2 MN [4 (PQf- (PT)2]

2yja2 - R2 sin2 A (4i?2 cos2 A — + a2)


and furthermore
4 R2
r= i?2 sin2 A ( 4 cos2 A — 12 +dA.
37ra2 i?

Substituting
sin A = a sin 8
/ ST
l
t 0 J P
5.12 Examples of the calculation of the dissipation function 257

FIGURE 5.17.
yields
7r/2
= ^ [. 3TT cos2 9 sin2 e
4 1 r2 de.
2
l-^sin *

We have

(5.12.15)
where K and E are the complete elliptic integrals of the first
and second kind with the modulus
0< k = — < 1.
a
In eqs. (14) and (15)

Vox +oVy,2
R= (5.12.16)

The limiting case = 0, which describes rotation about the


fixed axis passing through the circle centre O, is obtained
from eq. (14) by setting R = 0. In this case
E(0) = K (0) = ITT, r* = |a, <!> = ^fGa\uz\, (5.12.17)

which can be proved easily.


258 5. Work and potential energy

The second limiting case is the translatory motion.


Recalling the expansions

*<*> = f(1 + T + -)’ E


^ = Ul-T +
-

we obtain from eq. (15) by taking the

f { ^ [ ( 1 + 7*2) ( 1 _ T +2 " ) _
limit that (l+2fc + ...)(l + ^)]}

=1
= ^73 (1 + ™2)(1--/'2 k2—>0

and hence

<S> = fG(R\wz\)R_too = fGv0, (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(2 32 . .iiQ
V =-^r, * = —fGa\uz\ = l.l3fGv0,

where vo denotes velocity of the disc centre. This expression


is easy to derive directly from eq. (11) by setting r (A) = 2a cos
A and taking the lower and upper limits of the integral as —7r/2
and 7r/2, respectively.
The generalised forces of friction corresponding to the
quasi-velocities VQ and voy which are the projections of the
X

resultant of the frictional forces on axes Ox and Oy fixed in the


plate are, due to eqs. (11.4) and (16), equal to2
_ d$ d$dR v0x d$ dk
x
dvox dR dvox \uz \ vo dk2 OR ’
_ v0y d$ dk2
y
~ \UJz\ VO dk2 dR
Applying the relationships
dK IE - (1- k2) K dE 1K-E dtf ~ 2
(1 -k2)k2 ’ dk?~~2 k2

we obtain
p = 0, PUJ = |a/G signor,
(5.12.22)
b) translatory motion (R —► oo)
Q II

(5.12.23)

c) rotation without slip (R = a)

„ 8 fG „ 8 fG . P= 3n ' F" = 9„ (5.12.24)

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 $ for a polygonal plate is cumbersome but
elementary. Figure 5.18 shows a rectangular plate having
260 5. Work and potential energy

the simplest case, that is rotation of a fixed axis passing


through the centre of symmetry O, we obtain by eq. (12)
Ai A2
fb 3
a dX
3

12 ab J cos3 A ^ J cos3 A

where
a
\ b
tan AAi = — tan A 2 =
b a
Here similar to any polygonal plate the problem is reduced to
integration of expressions containing the following integrals
tan . 7r A
/ A
dX
cos3 A
+ ln
4 2
cos we obtain
tan| +

In the case under consideration


^/Gk |{vW&2+
$ (5.12.25)

+ a3 In 4. I 1n 1 A2
1 3
2
b In tan tan| +
T

ab

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 [68] we consider a system of aerodynamic forces
acting on a spinning shell moving through the air at rest. The
5.13 Aerodynamic resisting force 261

to the earth is described by the velocity vector v of the pole O


and the
angular velocity vector u. 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 m° on vectors v
and u which
is indifferent to the choice of the pole.
Projections of v and u on the shell axis are denoted by u 3
and a;3,
respectively. It is presumed that they are much greater than
the magnitudes
u* and UJ* of the components v* and CJ* lying in plane Oxy.
The vectors F and m° are also represented in the form
F = F* + kF3, m° = m? + k mg,

(15.13.1)
where the transversal force F* and the transversal moment
are respec-
tively the components of F and m° in the plane Oxy and k
denotes the
unit vector of axis Oz.
(15.13.3)
The axial force F3 and the axial moment m® are assumed to
have the
form
F3 = -pa2f3vl, m3 = -paig3v3uj3, (15.13.2)

where p denotes the air density, a the shell radius and / 3 and g
% are non-
262 5. Work and potential energy

Components 6ik x v* and C\U)* of the transversal force and —


a[v*and d!x k x u>* of the transversal moments characterise the
so-called Magnus effect which appears due to rotation of the
shell about axis Oz. The coefficients b\, ci, a'i, d!x are therefore
taken to be proportional to whereas the other coefficients are
taken to be proportional to v%. Thus we have
®i = fiV3, h = f2aw3, Cl = gia2u3, di = g2av 3 ,
a[ = f[auj3, b[ = f2v 3 , c[ = g[av3, d[ = g'2a2u)3.

The expressions for the transversal force and moment take


the form
F* = pa2 (-/1V3V* + f2au3k x v* + 0ia2w3a;* + g2av3k x a>*),
m? = pa3 (-/(aa)3v, - f2v3k x v* - + g2a2L03k x
u>*),
(5.13.
4)
fi, /2,91,92, f[, f2,g'i, g'2, /s, £3

(5.13.5)
appearing in formulae (2) and (4) depend on v^/c and auj^/c. The
signs are chosen so that the coefficients /1, g2, , #i 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 mc is determined by the relationship
m6 = m° + ek x F = m° + ek x F*,
(5.13.6)
where ek — C(5 denotes the position vector of point O with the
origin at the inertia centre C. Force F and moment mc should
now be expressed in terms of the velocity of the centre of
m6 = + rn^k = + ra^k + ek x F*,
F = F? + F3 k = F* + F3k,
c

from which it follows that eq. (2) and thus, the aerodynamic
functions fs
and gs remain unchanged under change of the
pole, i.e. (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
Ff = F* = pa2 [~fiv3 (vp + ek x u>*) +
f2cuv3k x (vp + ek x u>*) + gia2uj3uj* + g2av3k x u>*] ,

which is reduced to the form


F* = pa2 (-/f u3vp + /f aw3k x vp + ppa2 w3 u>* + g%av3k x u>*) ,
(5.13.
9)
where
fi = fi, f? = f2, 9? =9i--f2, 92 = 92 --fi- (5.13.10)
a a
We also have
mp = pa3 [-f[aoj3 (vp + ek x w*) - f2v3k x (vp + ek x w») - g[av3w* +
g2a2uj3k x w*] + po ek x
2

( ~ f i «3vp + /20 ^3 k x vp + a2 w3 u>* + g%av3k x w*)

and this expression can be simplified, to give


mp = pa3 (-/(caa;3vp - f!fv3k x vp - gfav3u>* + g!fa2w3k x u>*) ,
(5.13.1
1)
where

fiC = fi + ~hi 9ff = 9i-------------U2 ~ 92)-------%


a a a* > (5.13.12)
fF = f* + “/lj ^2 C ~ 92 ~ ~ { f i ~ 9i) ~ ^2/2- y

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 components 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 /2,<7I , / I ,#2 are set t° zer°5 then, as follows from
264 5. Work and potential energy

FIGURE 5.19.
assumption yields the following expressions for the force and
the moment
related to an arbitrary pole
F = pa2 (-kf3v% - (5.13.13)

fiv3v*+ g2av3k x u > „ ),


m° = -pa3 (kg3auj3v3 + f^k x v* + #(au3u;*),
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 6
with the velocity vector, the unit vector ii having the direction
of the transversal component v* of the velocity vector v and
the unit vector 12 = k x ii. Angle 6 is assumed to be small, i.e.
cos <5 ~ 1, sin <5 « 0, so that the velocity value can be
identified with v%. The vector a;* of the transverse angular
velocity of the shell lies in the plane of vectors ii, i 2-
The aerodynamic forces reduce to the forces R and L and
the moments M and H.
The force R represents the head resistance and is taken to
be equal to R = -pa2fRv3\ = -pa2 (fRkvl + fRv3v^),

(5.13.14)
5.13 Aerodynamic resisting force 265

where we neglected the term proportional to <5 2 and took


V
2
8\\ = vv*\\ « V3V*.
The tilting moment M and the retarding moment H are given
by
M = -pa3/MW3k x v„ H = -pa4fHv3u}*.

(5.13.16)
Expressions for the resultant force F and the resultant
moment mc of the aerodynamic forces about the centre of
inertia take the form
F = —pa2 [fRkvj + (fR + fL) v3v*] , 1 ^ 13 17^
mc = -pa (fMv3k x v*
3
+ fHav3a>*). J ’ ‘
These relationships enable one to link the aerodynamic
93=0, g<i= 0.
(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 x v in the
expression for F when the pole is chosen at the centre of
6
The fundamental equation of dynamics. Analytical statics

6.1 Lagrange’s equations of the first kind


are the forces exerted at the points in the system
Constraint forces
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 whereas that of the active
forces is denoted by F
Introducing the constraint forces enables us to write the
differential equation for any particle in the form of Newton’s
second law
raiW* = F * + R*, ( i = l , . . . , A 0 , (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 wi the acceleration
vector referring to an inertial coordinate basis Oxyz
Wi=Vi = 'fi. (6.1.2)

Projecting eq. (1) on the coordinate axes and applying the


notation of eq.
(1.2.1) we have
mvZj, = Fy + Ry {y = 1 , . . . ,3N), (6.1.3)

A. I. Lurie, Analytical Mechanics © Springer-


Verlag Berlin Heidelberg 2002
268 6. The fundamental equation of dynamics. Analytical statics

where m3s_2 = m3s_i = ra3s for s = 1 , . . . , N. In addition to these 3N


differential equations we have r equations for the holonomic
constraints
fk{Z1 „...,Z3 N;t) = 0 (k = l,...,r) (6.1.4)

and r' equations for the non-holonomic constraints


3N
5^k^„ + <fc = o (k = i,... y).

(6.1.5)
U=1
The unknown variables in eqs. (3)-(5) are the 3 N
coordinates , , . . . , £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.
N 3N
8'W = 6n =
(6.1.6)
2=1 V=1

The constraint equation holds also in the varied state, that is


along with eq. (4) the following equation holds, too
+ , • • • , £3iV + ^3JV > 0 — 0.
fk (£1
This means that the variation Sfk of function fk must vanish, i.e.

<A N
- g t^=!(tfe+i^+§H
N
= 53 gradi fk ’ 6r
i = 0 ( k = l , . . . , r ) .
(6.1.7)
2=1
Equalities (5) imposes additional r' conditions on variations
3N N N ~

S hv&iv = (akitixi + bki6yi + CkiSzi) = Y *ki • =0


2 =1 2=1 2 = 1 (6.1.8)
(fc = 1, . . . , 7*') .

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
6.1 Lagrange’s equations of the first kind 269

of the constraints and vectors e^, i.e.


r r'
R< = £ Afc grad;/fc + ^ A'feefei (i = l,...,N)
(6.1.9)
k=1 k=1

y>X (»• = 1............3JV). (6.1.10)


k=1 fc=1
Coefficients A& 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 -
3N| £Afc^ + £ ^klku
£^
for the constraint forces (10)ifc= into1 the
< elementary work ( 6) and
*" k=1
taking into U=1
account eqs. (7) and (8) we 3Nhave 3N
= £A*£|r^ + £ A * £ ^ ^ = o,
k=1 i/=l ^
k=1 v=l
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 Afc and Xfk so that the 3N quantities Rv
are represented by linear forms (10). Indeed, multiplying each
of equalities (7) and (8) by —Xk and —A^., respectively, and
summing up the products obtained with eq. ( 6) we arrive at
the equality 3N 3N 3N
£ - £ A f e £ ?L-6(: - £ A ' £ lkv6^
V

k= 1 v=l ^
U=1 k—1 u=l
3N A
=£ ^-X>lr-£ ^ 6L = 0. (6.1.11)
k=1 ^ k=1
V=1

We take now the r + rf coefficients \k and \rk 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

Rjy — ^ ^ Xk dfk + A khu {v= 1 , . . . , r + r ) . (6.1.12)


k=1 dL k=1
270 6. The fundamental equation of dynamics. Analytical statics

By virtue of the general assumptions about the constraints,


these equations can be resolved for A*, and A^.
Equality (11) now takes the form
3N
£ Ru = o, (6.1.13)
i/=r+r' k=1
+1 fc=l
with 3N — r — rf variations being independent of each other. For
this reason, the coefficient of each of these variations in eq.
(13) must vanish which yields
Rv =
+
'Yh'K'klkv {v = r + r' + 1, . . . ,31V). (6.1.14)
fc=i k=1

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 Xk and \'k which results in
the following equations of motion for the particles of system
A
l„ = Fv + Y< *|T + EA^ (^ = 1,. - - ,31V)
mv (6.1.15)
k=1 k=1
or in vectorial form
r r
miWi = rriiVi = Fj + ^ Afe grad; fk + ^ A'fcefci.
(6.1.16)
k=1 k =1
These equations are called Lagrange’s equations of the first kind. The
problem is determinate since the 3 N equations (15), along
with r + r' constraint equations (4) and (5), have the same
number of unknown variables
^1 5 • • • 5 Vj Ai,...,Aj./.
£I> - * * > £ 3 NI
In what follows we will return repeatedly to the theorem
proved here: the sufficient condition for the following equation
n
Y^X 6X = S S 0,
(6.1.17)
S —1
n
^afes&rs = 0 (k = l,...,r) (6.1.18)
S=1
6.2 Ideal constraints 271

is that functions Xs are represented in terms of the r


independent Lagrange’s multipliers A i , . . . , A r
r
Xs = 'Y^\kaks (s = l, . . . , n ) .

(6.1.19)
k=1
Conversely, if eq. (17) holds, one can always find the r
coefficients A& that represent functions Xs by linear

6.2 Ideal constraints


Holonomic and non-holonomic constraints satisfying the
condition under which the elementary work of the constraint
forces for any virtual displacement 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 Sr of the point of
contact of the body and the surface. The elementary work R •
Sr 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 • Sr = 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
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
J2n^ = J2Ri'6ri = 0- (6.2.1)
v=l i =1

Replacing here <Sri by means of eq. (1.6.7) or eq. (1.6.14)


in terms of variations of the generalised coordinates or
variations of the quasi-coordinates, respectively, we obtain
n N p. n N p.
j T R i - ^ = 0, J^Sns^TRi F
0. (6.2.2)
S— 1 i= 1 s
s =l i= 1 dir*

By definition (5.1.3) and (5.1.7) the internal sums in these


equalities are the generalised forces for the constraint forces
a = X>-|p p‘*s = f > - - & = 0 ’
1=1 V i (6.2.3)
dirs
=1

and eq. (2) takes the form


N N
E^=0’ J2
p
:Sns=0. (6.2.4)
S=1 s= 1
In particular, if gi, ...,g n are independent generalised
coordinates and the non-holonomic constraints are absent,
then variations 6qs as well as variations of the quasi-
coordinates <$7rs are independent. It follows from eq. (4) that
all of the generalised constraint forces are equal to zero
Q* = 0 or Pfl*= 0 (s = l , . . . , n ). (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)
rmwi = Fi + Hi (i = 1,.. . , TV) .
(6.3.1)
The fundamental equation of dynamics is derived by means of
eliminating the constraint forces from the above equation. The
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 constraints.
Elimination of the constraint forces is achieved with ease in
the case of ideal constraints. Recalling definition ( 2.1) of the
ideal constraints it suffices to multiply each equation in ( 1) by
<5r* and sum up the results, to get
N N
^rriiWi ■ 6 ^ = • <5r*.
(6.3.2)
i=1 i=1
Lagrange referred to this equation as the fundamental equation of
dynamics. 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
n-f-ra
= E 7T6^ (6.3.3)

and the fundamental equation reduces to the


form
Nm w n-\-m r\ N n-f-ra ^
E * * • E = X> • E
2=1 S= 1 ^ 2=1 S— 1 ^

Introducing the generalised force we have


n+ra / N \
E (Em*w* • - Qs I = o. (6.3.4)

Variations 6qs are related by m + r' equalities


n-f-ra
J2 ^±6 q = 0 (k = l,...,m),
s=l ^s
s
(6.3.5)

n-f-ra
j2aksSqa = 0 (k = 1,... ,r'),

(6.3.6)
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
m w
dr m
dF r

2 2 • ~r,— — Qs F ^kaks (5 = 1, . . . , n + m),


2=1 k=1 qs
qs
k=1
(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 #i,..., gn_|_m, with m being redundant.
In the case of holonomic system and absence of redundant
coordinates eq. (7) simplifies and takes the form
N
fir
Y]rniwi--^±=Qs (s = 1,... , n). (6.3.8)

The sum
V*. dp
k
^ lw,
k=1 Hs

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

^k ks•
a

k=1

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 dynamics known as Lagrange’s central equation.
We have
miWi • <Sri = miWi • <Sr* = —rn;v* • <Sr* - • ((Sr*)*
at
= ‘ ~ miVi ’ ’ [^v* “ (*,)•]
The last term vanishes if the operations of varying and
differentiation are interchangeable. Noticing that
Vi • <5vi = (Vi • Vi) = 1^1,

we reduce the fundamental equation of dynamics ( 2) to the form


,N N
1miv N
—Y2 mivi ■ =5-^3 i y^,Fi'
+

2=1 2=1 2=1


6.4 Rearrangement of Lagrange’s central equation 275

or
7 ^
— YmM • 6n = ST + S'W, (6.3.10)
2 = 1

where T and <5T denote respectively the kinetic energy and


its variation and S'W is the elementary work of the active
forces. This equation is termed as Lagrange’s central equation.
When the operations of varying and differentiation are not
interchangeable, i.e. the law (1.7.5) does not hold, we obtain
the following equation
7N N
Y171^ ■ ^ *
r =
ST + S'W + Ymivi • [(&»)* - <5 »] >
V

(6.3.11)
2=1 2=1

6.4 Rearrangement of Lagrange’s central equation

Consistent with eq. (5.1.1) the following expression


Nm v
'Y i i-8ri (6.4.1)
= 2 1

can be treated as the elementary work of the linear momenta


(also
known as impulses) due to the virtual displacements of the
system particles. Similar to the generalised forces Qi we can
introduce the generalised momenta. They are expressed in
terms of the momenta in a manner like the generalised forces
are expressed in terms of forces F*, i.e. by means of eq.
(5.1.3)
N ^
Ps = YmiVi ■ 7T~-

(6.4.2)
U dq
°
The equality
corresponds then to relationship (5.1.4). Using
transformations (1.3.5) we can recast eq. (2) in the form
N dri ___dvi _ d 1
ps = Y m*Vi ~dqr ~ Z^miVi ' ^7 “ miWi (6.4.4)
2=1 2=1
dqs dqs 2 2=1
276 6. The fundamental equation of dynamics. Analytical

statics
or
dT
PS = T r r - ( s = l , .. . ,n ) . (6.4.5)
dqs
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
in n
jtJ2p^ = 6T+'E^6^- (6-4-6)
S=1 S=1
When the forces are potential then due to eq. (5.3.9)
ST + S'W = 6T-6n = 6(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 ( q i , . . . , q n , q i , . . . , q n ; 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
ftJ2PssQs = SL. (6.4.9)
S =1
When the rule ” dS = Sd” does not hold the right hand sides of
eqs. (6) or
(9) should be completed by the term
N N n
f)
^ 2 rtiiVi • [(<5rj)* - <5vj] = 'Y^mivi ■ ^ [(6 qs)* - 6 qs]
i=1 i=l 8=1 °^s

n N n

= y2 i im v
• -f 1
= T, [ ( ^ » ) #
ps _
^s] • (6.4.10)
8—1 j=l ^ S=1
Here formulae (1.7.3) and (2) are used. The central
fundamental equation
is then written as follows
,n n n
jtY,p* (i* = 6 T + J 2 Q ' 6 * * + -
8
*«•] • (6-4-n)
S=1 S=1 S=1
This equation should be expressed in terms quasi-
velocities and varia- (6.4.12)
tions of the quasi-coordinates. Repeating transformation
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) nbsrPr
P* = J2 S E6' d T dT
( s = 1, . . . , n ) . (6.4.13)
r= 1 r=l d q r dojs
The quantities p* which are the derivatives of the kinetic
energy with respect 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 dT

OUJ\
- = ©11^1 + ©12^2 + ©13^3 = ,

dT (6.4.14)
— — + ©22^2 + ©23^3 =
Q 2\ U ) \

>
dujo
dT

as follows from expressions (4.7.4) and (4.8.13). Therefore,
- = ©31^1 + ©32^2 + ©33^3 = ,

the momenta corresponding to the projections of the angular


velocity are the projections of the resultant angular
momentum about this fixed point on the corresponding axes.
This illustrates the importance of the quantities p* in
mechanics.
The inverse relationship to eq. (13) is
Ps = y~] arsP* (s = 1,... , n).

(6.4.15)
r—1
Thus making use of formulae (1.8.4) we have
n
S= 1 n n

T,Psr= [(^s)*
1

6<
is} = Tpr E t= 1 1=1
ars
[(^»)* - <%]
1=1

and the central equation (11) takes the form


= ST + ±Ps6ns + ±^-[(S,sr-S.s] (6.4.17)
S= 1 '

s= 1 s=l
nnn
s=l r=1 s=1
r=1 t=1
If the active forces are potential forces, the right hand side
can be cast as follows
n
ST + ^2 s^s = 6 (T - n) = SL.
p (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
dT _ dL „ dT dL (6.4.19)
dqs dqs ’ Ps du)s du>s

6.5 Equilibrium of the system of particles


When the system is in equilibrium, the acceleration w* 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)
F* + R;=0 (i = 1,... ,A0-

(6.5.1)

Using eq. (1.15) we can write 3N equations of equilibrium

Fv + J2xk^ + J2Klk, = 0 (^ = 1, . . . ,37V).


(6.5.2)
fc=l k= 1

Of course, in assuming an equilibrium we assume that the


constraint equations (1.4) and (1.5) do not contain time
explicitly and moreover the free terms gk 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 4 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
equilibrium when all 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 4- r 4- r' quantities remain
indeterminate.
n+ra
^ ^ &ksQs — 0 (A = 1 , . . . , r ) ,
s= 1
J
(6.5.4)
6.5 Equilibrium of the system of particles 279

we obtain by setting in the left hand side of eq. (3.7) w* =0,


the n + m equations of equilibrium
r
QJ? ^
QS + ^k~g~~ + ^kaks = 0 (k = 1, . . . , n + m).
(6.5.5)
k=1 k=1

These express the condition that the sum of the prescribed


generalised forces and the generalised constraint forces
BF r' x a
171

Q*s = J2 ~&r + ^2 'k ks- Xk


(6.5.6)
k=1 qs
k=1

vanishes. The problem is defined when non-holonomic


constraints are absent. The first sum on the right hand side of
eq. (6) can be written in the form

_d y A,
f>^=
_ ^ dqs (6.5.7)
k=1 dq k=1
since, by virtue of eq. (3), the following sum
dx
Y' k F k=1

can be cancelled out.


Let us consider the case of potential forces, then
Qs =
(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

d II + AkFk j — 0 (s — 1 , . . .
dq n (6.5.9)
s k=i

These are extremum conditions of the function


n ((/l, • • • , Qn+m)
subject to m constraints (3).
In the case of no redundant coordinates and non-holonomic
constraints, the equilibrium equations
Qs = 0 (s = 1, . . . ,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
fi=0 (s = 1,... , n ),

(6.5.11)
dqs
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
qi+6q!,... , q^ + 6qn
the increment in the potential energy
An = n («? + s qi , + 6q n) - n (<??, (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 All 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
the equilibrium position is stable if the potential
Lagrange and Dirichlef
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]. (6.5.14)
We formulated here the problem of equilibrium by using the
principle
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 become as simple as possible
= qn+1 ~ qn+1 = 0,... , Fm = qn+m ~~ qn+m = where q^+k
are
constant values of the redundant coordinates. Then
dFk _ £ _ J 0, n + k^s,
dqa~°n+k'a~\l, n + k = s,
and the equilibrium equations take the form
m =
Qs T ^ ^ &s,n+k^k 0
k=1
or, more specifically,
Ql — 0, .. ., Qn — 0, Ai — Qn+l, • • • 5 Am
— — —Qn+m- (6.5.16)
For example, let us consider the equilibrium of a free
particle on a surface. Entering the curvilinear coordinates
g 1 , # 2 , # 3 , such that q3 = q$ corresponds 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
6'W = F • Sr = F • rs6q ,
s
where rs denotes the base vectors of the surface. The
equilibrium equation
(14) yields

F • ri = 0, F • r2 = 0, Ai = —F • r3.
These equalities hold on the surface, i.e. at q3 = qfj. Expressing
the vector of the active force in terms of its covariant
components
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 i?3 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, BC attached to
each other
by joints A and B, see Fig. 6.1. The joint O is fixed and the
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
respectively. The
thread
tensions TI,T2,T3 are required.
The active forces are the rod weights GI,C?2,C?3- The
II — — ( G \X\ + G 2X2 + G3 X3 ),
where X\,X2-)xz denote the coordinates of the centre of gravity
of the respective rod

Xi = Si COS X 2 — ^1 COS T ^2 COS X 3 = 1 1 COS ( p 1 “h I 2


COS </?2 H- 53 COS ( f . 3

Here luhih denote the rod lengths and S I , S 2> S 3 the distances
between their centres of gravity and joints O, 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 , <^2 ? ^3- These three constraint
equations can be cast in the form of eq. (5.15)
— =
Vi ~ Vi = ^2 ^2 0, ^3 —
^3 =
d-
We obtain then that
n= - (GiS! + G 2I 1 + G3/1) COS^! - (G 2S2 + G 3I 2) COS - G 3 S3 COS(p3
6.6 Examples of deriving equilibrium equations and constraint forces 283

and by virtue of eq. (5.16) we obtain the constraint multipliers


Ai — (GiSi + G2I1 + G3Z1) sin^J,
A2 = (G2S2 + G3Z2) sin(^2, (6.6.1)

>
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
T\6yA + T261JB + TsSyc = Ai 6(p1 + A2 6<p2 + A3 6(p3. (6.6.2)
Since

by A = l\ cos (p^6(pi,
dyB = h cos tpifitpi + I2 cos ^28^21
$yc = h cos ifi6cp1 + Z2 cos (p2d(f2 + h cos (p3d(p3
we equate the coefficients in eq. (2) for the independent
variations 6(ps, to obtain a system of three equations
Ai = (Ti + T2 + T3) li cos^J, A2 = (T2+T3)Z2 cos (^2, A3 = T3l3 cos
<^§, where (ps was replaced by (p®. Solving this system for Ts
yields

It is easy to prove that the system is in equilibrium since the


resultant moment of the gravity forces Gs and the thread
tensions Ts about the fixed joint O is zero.

6.6.2 Equilibrium of a heavy rod gliding by their ends on a


smooth surface
Rod G of length l 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

* = f(x,y) • (6.6.4)
-Pi ~Qi 1 0 0 0
0 0 0 ~P2 —Q2 1
X\ - x2 y\ -V2 Zl - Z2 X2 ~ X\ V2 - yi ^2 - z1
-Qi
1 0
0 0 -P2
yi - y 2 Zl - Z2 x2 - X\

The determinant of the following 3x3 submatrix

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 m, see Fig. 6.2, and the directions cosines of
the normal relative to the coordinate axes as a, /?, 7. Then
a =____, g8 =2_, -y — 1

y/1 +P +
2
q ’ y/1 +P + q ’
2 2
y/1 + p2 + <j>2
(6.6.6
)
6.6 Examples of deriving equilibrium equations and constraint forces 285

Let e denote the unit vector directed along the rod from A to B
and ei, e , e% the cosines of the angles between e and the
2

coordinate axes
x - X\ V2 ~ V\ l 5 ^2 ~ Z! (6.6.7)
e
i 5 e2
2
e3
l I'
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

n = ±G(z1 + z2).
Using the notation of eqs. (6) and (7), the equilibrium
equations (5.9) are
Oi 1A 1 — 2A 3/ e i 71 ? aX
2 2 — — 2A 3/ e i 72 ,

/?iAi = 2A3/e27i> /?2^2 = —


2A3/e272? (6.6.8)
Ai = 2Xsles + —G, A2 = —2X^le^ + —G.
Resolving them for Ai, A2, A3 yields
2A 3l = = \ G-
«i
PI

2 ei7x - ezai 2 e2^l - ez(3l


a2 \ G- P2 (6.6.9)
= --G
2 ei72 - e3a2 2 ei72 - e3/?2

7i = jGe2 71
Ai = \Gex e
27l ~ 6 3/?l
6l7l
e
“ ’ (6.6.10)
3<^l
A 2 = \Gex 72^a2 72
ei72 —
6272 “ 63/?2
’ form
The latter two equalities reduce to the
e
i _ ai _ a2 (6.6.11)
e
2 Pi P2
This relationship also satisfies two of the equalities (9). It
expresses the fact that the normal vectors mi and m 2 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 mi and m2. We have
exk
n=
ex
and thus k|
1
..1
n rm = (nij x( e) • k (aie2~Plei) = 0 (i = 1,2),
|e x k| = e x k|
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 e x ntii |e
x rrii|

and thus
e-21 i~ e3ai - eifc - e2ai
nI —j-------------j—, n 2 = —j-----------------------j—, ns = —;-------------------j—
|e x nii| |e x nii| |e x rrii|

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
n\.
Another relationship which is a consequence
oq a2 ei
of eq. (9) is
£3 /?! @2
Ti 72 +
or Ti
=2

e
, 72
^_
(6.6.12)
= 2

2
Its meaning will be explained in what follows.
Therefore, we obtained three equations, eqs. (11) and (12),
and the equilibrium positions exist provided that these
equations along with the constraint equations (5), have
solutions.
Now we proceed to determine the constraint forces. The
resultant of R* 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 = Ai gradx Fx + A3 gradx F3, R2 = A2 grad2 F2 + A3 grad2 F3,
as F\ and F2 depend only on xi,yi,z\ and x2, y2l 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 = —T2. We obtain for i = 1, 2
1 Ge\ot\ ^ _ 1 Ge2pi
Nix
2 ei7j - e3at ’ ty 2 ~ e3’ Niz —Ai

and
T2 = -Ti = 2A le.
3

(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

z— c+c 1 - x1 +y 2

Dealing with the surface of revolution one can take the


plane Ozx as the plane in which the rod lies. It suffices to put y\
= y2 = 0 in the above formulae. The equations for determining
the unknown variables X\ and X2 are
7l 72 = o ^2 - *1 (x2 - Xi f + (z2 - Zif -l2 = o,
a\ OL2 X2 — X\ ’
where

z= c+c

Since
7 la2

a p cx
we put x = a sin u in the above equations, to have
c2 COS U2 — COS U\
cot u\ + cot ^2 = z -------:-----,
a sin v>2 — sin u\

2 2
c l
(sin^2 — sinui) H—^ (cos^2 — cos^i) = —.
Introducing the half-sum and the half-difference of the angles
u\ and U2
U2+UX U2~ Ui

we can rearrange the latter equations


4a2’
sin a (cos2 (3 e cos
22
a
) = 0, sin2 (3 (l e2 cos2 )a =

where
c2 — a2
e2 =

Consider an elongated ellipsoid of revolution, i.e. c > a, 0 <


e2 < 1.
The equilibrium position a = 0 is not feasible as a > 0, thus the
feasible
288 6. The fundamental equation of dynamics. Analytical statics

corresponds to the horizontal equilibrium position A\B\ in which


angle (3 is defined by the following equation
sin2/ ? ( l - e 2 ) =
and the problem has a solution (which implies that an
equilibrium position exists) when the obvious condition
l< 2a\/l - e2

holds.
The second equilibrium position takes place when cos 2 (3 =
e cos2 a, which is equivalent to l = 2a sin2 (3. It can be proved
2

that the rod passes through the focal point of the ellipse. A
simple geometrical proof is suggested in [101].
If a flat ellipsoid (c > a, 0 < e 2 < 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 orientation of the trihedron of
central axes Gxyz (see Fig. 5.4a) relative to axes of fixed
directions the potential energy takes stationary
values.
These parameters are Euler’s angles which appear in the
potential
6.6 Examples of deriving equilibrium equations and constraint forces 289

According to eq. (5.11) the required positions of trihedron


are determined by the conditions
Gxyz

on _ on n an m
mN = =
°’
TO3
'777 = °> 3 = —sr = °-
dip dip
By means of formulae (5.2.18) this can be cast as follows
mi cos cp — m2 sin cp = 0,
(mi sin cp + rri2 cos cp) + m3 cos $ = 0,
m3 = 0,
where m i , m 3, m 3 are determined by eq. (5.5.10). Thus there
exist two possibilities
a) mi =0, m2 — 0, m3 = 0;
b) m3 = 0, sin^^O, mi cos<^ — m2 s i n < ^ = 0.

Case a) occurs under the conditions


/?7 = 0, 7a = 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 directed along M(J. For instance, if a = (3 = 0, |
y| = 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 M&.
In case b) we direct axis M( along M(^, then £G = yG = 0 in eq.
(5.5.11). As sin 1? = 0 we obtain
a = (3 = 0, 7 = 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 = 1 [(2 - £! - £2) a2 + (2£l -e2-l)I32 + (2e2 - e, - 1) 72] (6.6.15)
only in an additive constant and an immaterial positive factor.
In accordance 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
2 — £i — £ > 0, 2S2 ~ £l ~ 1 < 0.
2
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
2ei — £2 — 1 — 0
shown in Fig. 6.4 splits the plane of values £1,62 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 $ = a2 + /?2 + 72 - 1 = 0 is sought from the equations
±(F-\*) = 0, A(F_a4,) = 0 , ± { F - A*) = 0

which are recast as follows


(2 — £\ — £2 — A) o = 0, (2ei — £2 ~' 1 —■ A) f3 = 0, (2^2 ~ £1 —1—
A) 'y = 0.
(6.6.1
6)
Since a, /?, 7 can not be zero simultaneously there are three
cases
a) a = f3 =
0, A = 2^2 - ^1 - 1 < 0,
b) (3 = 7 =
0, A = 2 - £1 - £2 > 0,
7 = 0 = 0, A = 2E\ — £2 — 10.
c)

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 a,/?,7, respectively, and summing up the products
yields
F = A (a2 + /32 + 7 ) = A.
2

Therefore, case a) corresponds to the minimum value of the


6.6 Examples of deriving equilibrium equations and constraint forces 291

of Lagrange-Dirichlet’s theorem case a) describes a stable


equilibrium position in which the ’’long” axis of the inertia
ellipsoid lies along M & . Case b) corresponds to the
coincidence of the directions of the ’’short” axis and M & 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 Mi,..., M m . The other ends of the rods are fixed at
immovable points Si,..., S 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 Fi,..., Fn. 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 O x y z denote the system of axes fixed in the body, V
and m° denote the resultant force and the resultant moment
about the pole O of the system of forces Fi,..., Fn, respectively,
and denote the position vectors O M k of joints Mfc. 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 of
the joints M k are different and are given by
uk = U o T 0 rfc.
X

and thus
fk=u0 • e
fcefc • U0 + 0 • (eg x rfe) (eg x rfe) • 9 - 2u0 • eg (eg x rfe) • 9. The potential
energy of the elastic rods can be written in the form

n u0 • ^2 Cfce°e° ■ u°+ (6.6.18)


fc=1 k= 1
m
6
■ X] (e°
(e° X rfe) • 0 - 2u0 • ^2 ckek {el X fe) • 0
Ck x rfc
) r

k—1 fc=l
This is a quadratic form of the projections of the vectors Uo
and 0. Within
the accuracy assumed it makes no difference whether the
axes are fixed in
Uo A B u 0 V
e B 'C e m°

where A , B , C denote the following matrices


m m m
^ = Y1 c
B
* e ° e fc> =
c e
k kr'k4, c = - Y2 ck^kr'ke%.
k= 1 k—1 k= 1

Determination of the column-matrices U Q and 0 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 |Q| 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 constraint forces, [71]. The
resultant vector and the resultant moment about the pole O
are denoted by VQ and m^, respectively. The rod elongations
293 forces
6.6 Examples of deriving equilibrium equations and constraint

FIGURE 6.5.

from their initial states are 6k . We consider passage from the


initial equilibrium configuration S o to configuration S due to
a new resultant vector V and the resultant moment m° about
the same pole O . In other words, we look for the
displacement vector UQ of the pole O and the rotation vector 6
describing this passage.
As shown in Sec. 5.8 the potential energy of the k—th rod in
configuration S is as follows
2°k (fk + 25*;/*; + S2k) .

Thus, the elongation should be found up to the values of


second order. Taking the rotation vector 6 to be small, it is
necessary to retain in Rodrigues’s formula (3.1.11), for
displacement of point M°, only the terms of second order in 0.
Then we obtain
uk = u0 + 6 x rk + ^0 x (0 x rfc) = u'fc + Auk.
Figure 6.5 shows that
ll
2 (l + fk) — l u + 2/^e^ • u*;,
k k k 2Zjjl/fc + f — k + ^k k * ufc-
2 u e

k
The solution of this equation is sought in the form
fk = u'fc • e°k + £fc,

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
u* 4Y + Au'fc •ek'o
294 6. The fundamental equation of dynamics. Analytical statics

An expression for 6k fk calculated within the above accuracy is


[uo -eg + 0- (rfe x eg)] + ^ ||u0 x eg|2 + |eg x (0 x rfc)|2
hfk = &k
-2 (UQ x eg) • [eg x (0 x rfc)] + Zg (eg x 0) ■ (0 x rfc)} . (6.6.20)
Let us adopt that the forces Fj, ...F° impressed on the initial
configuration 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
An2 = -v0 • u0 - m? • 0 + i* • Q° • 0,

where tensor Q° is given by eq. (5.3.24).


The first term in the expression for the potential energy of the
rod system
'm m m

n=2E+E+2E CkS
l
k=1 k=1 k—1
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 /|. The second term is denoted by Alii. 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 SQ. We now have
Alii + An2 = E cfc<5fcefc • (uo + 0 X rfc) - (V0 • uo + m? • +
fc=i
xeg|2 + |eg x (0xrfe)|2-
\ E T [| U O
l
l
2 (u0 x eg) • [eg x (9 x rfc)] + Zg (eg x 9) ■ (6 x rfc) + ^0 • Q° •
6 .
(6.6.21)
As So is an equilibrium configuration the variation of the
potential energy at UQ = 0 and 0 = 0 must be equal to zero.
This variation is
(OT)«o=O,0=O = ( - v0 ) • <$uo + ( Ec*«*r* x e° - m0 ) •
\k=1 Kk=1

In view of the arbitrariness of variations <5 UQ and <50, the


condition under which this variation vanishes, yields the
following equation
E - Vo = 0, E rfe x CkSke°k - mo = 0. (6.6.22)
fc=l k=1
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 expression for the potential energy vanish. Now using
eqs. (18) and (21) we arrive at the expression for the potential
energy

n = i [u'0 (A + A1) u0 + 2u'0 (B + BO0 + O' (C + Ci) 6} ,

(6.6.23)
B. = -E^(£

- 44 ')
ii
k—1 k—1 E r
Ck^ E - 4 ' k) + >
-E
k ~Tki - 44') ^ + ECk6k
Cl
(44
m k=1 il k=1
E (EF?4 - F^) ■ (6.6.24)

Here and denote the position vector of the joints mounted on


the body and of the points of application of forces in the
equilibrium configuration
So, and are the unit vector of the rod SfcM^ and its length in
this configuration, fk denotes a skew-symmetric 3x3 matrix
accompanying vector Yk 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
i in
-^(c^+^l + Anj

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 + A\ B + B\ (6.6.25)
Bf + B[ C + Ci

can hold for certain initial elongations 6k 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 determinant \M\ =0. Then the system of
296 6. The fundamental equation of dynamics. Analytical statics

no solution for arbitrary V and m° and there are no equilibrium


configurations S close to So- 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 m° 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 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 <$ro of the virtual displacement of the pole
O of the axes Oxyz and by vector 0 = 0zi3 of the infinitesimal
rotation about axis Oz. The elementary work of the active force
F in this virtual displacement is
6'W1=F-6TO+m°0z,
where m° denotes the moment of force F about axis Oz.
The value of the frictional force on the elementary surface
do of the contact of the plate and the plane surface is equal to
fpdo, / 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 equilibrium is broken.
The elementary work of the frictional forces due to two contact
surfaces S is then given by the following expression

8'W2 = -2 JJ
s
fp^ • (<Sr0 + i30z x r) do,

where r denotes the position vector of surface do. Introducing


the position vector rp of the instantaneous velocity centre P
one can represent the velocity vector v and its value v as
follows
v = 13UJZ x (r - r p ) , V = ICJ (r - rP)|.
Z

Denoting e = sign UJ we have


Z

S'W2 = -2e8r0 ■ Jj fp13 do - 2e6z J f *>■


6.6 Examples of deriving equilibrium equations and constraint forces 297

Equating the coefficients of independent variations <5ro and


0Z to zero yields two equations

F = 2e Jj fp -3 * ^ do, m° =2e JJ f'P * rP \


d0
’ (6'6'26)
5 S

where the latter equality can be rearranged into the form

raf = 2erP • JJ fp+ 2e JJ fp\r - rp\do. (6.6.27)


5 5
Because
i3rP- (r - rP) = rP x [i3 x (r - r P ) ] ,
we obtain, by means of eq. (26) for F,

• II
fl
2d3rp ’W^\ do=rrxF
-
S

Let r7 denote the position vector of a point on the action line of


force F, then
i3m° = r' x F, i3mf = (r' - rP) x F,
where mf 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

2e
II M r — rp\ do. (6.6.28)

Expressions for the projections of force F on coordinate axes


is now needed. As
ii • [h x (r - rP)] = - (y - yp), h- p3 x (r - rP)] = x - xP
and
l r - rP \ = \J{x- x P ) + (y - yp f,
we obtain

f
-= II ' ’
,f1

F, = II fp |^j* = II f, |r - rP| do. > (6.6.29)


s s
298 6. The fundamental equation of dynamics. Analytical statics

Thus, introducing the function

V (xP,yP) = 2 JJ fp |r - rP| do = 2 JJ fp\j (x - x ) P + (y - yPfdo,


s s
(6.6.3
0)
we can reexpress the force F and its moment raf as follows

Fx = F
v = ~£Q^’ m* = £^ (Xp»VP) ■ (6.6.31)

Assuming a distribution of normal pressure p(x, y) and


calculating ^ 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 instantaneous 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
UJz. 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 uz coincide as > 0.
Let us determine under what condition the initial
displacement of the plate will be pure translation. In this case
S'W2 -2
— • (<5r0 + hOz x r) do
vo
s
r do x Vo • 13
^0

and the principle of virtual work yields


s
r do
yo • 13- (6.6.32)
^0

If we introduce the centre of pressure, which is the point


determined by the following position vector
fffprdo
rc =
JT7^’ m 33)
5
then the moment m® can be written in the
form
m° = (rc x F) • i3. (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 pressure.
Therefore, an initial pure translation in the force direction
takes place when the force reaches the maximum value
F 2
™ = JJ fpdo,

(6.6.35)
S'
at which equilibrium is still possible and passes through the
centre of pressure.
dxp dyp
= 0, = 0, (6.6.36)

that is, has a stationary value. In order to determine the sign of


the second variation b2"^ which coincides with the increment in
d/ when xp.yp are replaced by xp + 6xp,yp + 6yp
l/a 2
d-f2 0 d6x2V f . 2 d2^
dy2P Hi
C2T

6 ^ o ( »~ 2P + a---------3- -S PHP +
= X

2 \dx p dxpdyp
=^ JJ (y - ypf Sxp -2(x- Xp) (y - yp) 6xP6yP+ s

(x-xP)2Sy% JJ £j[(y-yp) x 6 P - (x - xp)6yP]2 do.

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 <S 2\I>
> 0 and the above stationary value is a minimum. Zhukovsky
referred to the point (xp, y*P) defined by condition (36) as the
frictional pole. When the body is subject to the pair of forces

mf = e* (x*P, y*P) = etf min ,


the initial rotation about the frictional pole occurs. No
equilibrium is feasible if the absolute value of the moment
exceeds ^min- As function ^ 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
*(!) K!)
R R
a a m\
0 0.6667 0 1 1.1318 0.8488
0.173 0.6817 0.1730 1.015 1.1449 0.8550
6
0.342 4
1.064
0.7247 0.3369 1.1870 0.8714
0
0.500 2
1.154
0.7897 0.4838 1.2669 0.8942
0
0.642 0.8677 0.6076 7
1.305 1.4038 0.9199
8
0.766 0.9487 0.7047 4
1.555 1.6375 0.9452
0
0.866 1.0227 0.7744 7
2 2.0631 0.9677
0
0.939 1.0814 0.8188 2.923 2.9665 0.9851
7
0.984 1.1189 0.8420 8
5.758 5.7797 0.9962
8 8 oo 1
oo

R
The table of functions aip and ip' R
a
Of course, the centre of the circle O is both the centre of
pressure and
the frictional pole. The latter statement is due to ip' (0) = 0 and
is also
clear from the symmetry of the problem. The initial translatory
motion
takes place under a force having the value = 2fG passing
through
the centre of the plate whereas the initial rotation about the
circle centre
occurs under the pair of forces with the moment F^aip (0) =
0.667Fma. In
the general case we can adopt that, due to the problem
6.6 Examples of deriving equilibrium equations and constraint forces 301

Since 0 < ^' < 1 the equilibrium is possible only if F < Fm. The
action line intersects axis Oy at the point
R\ 1
R aJ
V= a RN
a.

aI

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 obtained easily with the help of
Lagrange’s central equation. The equations will be derived
twice here. The first derivation will assume that the operations
d and 6 are not interchangeable, while the second one will
assume that the operations are interchangeable. In the first
case, i.e. if d6 ^ <5<i, 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

We also have

(7.1.2)

A. I. Lurie, Analytical Mechanics © Springer-


Verlag Berlin Heidelberg 2002
304 7. Lagrange’s differential equations

and substitution into (6.4.11) yields


n
E d dT dqs+Ps {Sqs)
dt + Qstq*
s =1
dqs
n
+ ^ jp8 (6q s )' - Ps6q s ■
s=i L

The underlined terms cancel out and we arrive at the equality


dT
-^--Qs)6qs = 0. (7.1.3)

Let us consider now the case in which dS = 6d. We now use


Lagrange’s central equation in form (6.4.6) which, after
repeating the above derivation, takes the form
n
d dT Sqs + Ps (Sq )'
E a + QsSqs.
dt s =1

dqs
The underlined terms cancel out since (dti)* = (8d)* 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 6 plays no principal part for
the derivation of the equations. This interchange simplifies
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 6qs are independent and thus the
coefficients in dfront
dT _of each 6qs in eq. (3) must be(7.1.4) zero
{s = 1, . . . , n ) .
dT_ dt dqs
dqs equations of the second kind. Their number is equal
These are Lagrange’s
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 relationships
n
^aksqs + ak = ° (k=l,...,l), (7.1.5)
S= 1
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)

"77 7T: Y— = Qs + (<9 = 1, ...,/) •


---------(7.1.6)
dt dqs dqs ^

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 -\-1 and
exceeds the number of degrees of freedom n — l which is the
difference between the number of independent parameters
qi,...,qn describing the system configuration and the number of
equations of non-holonomic constraints. Equations (5) and ( 6)
have n +1 unknown variables consisting of n generalised
coordinates qi,...,qn and l constraint multipliers A i , . . . , 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
dr i d d dri
dvi dt
dqs dt dqs dqs
However, due to eq. (1.3.11), we
have
d dvi dvi
Thus
drj d dvi dwi
niiWi dqs dt dqs dqs
d d1 d1
=
37 • v* - Y~omiVi ' v*
dt dqs 2 dqs 2
and
dr i d d v? u~ON u£ 2

dqs dt dqs ^ 2 dqs ^ 2


2—1 2—1 1=1
Taking into account the definition of the kinetic energy (4.1.1)
we obtain
N dr£ d dT _ dT (7.1.7)
2=1 dq dt dqs dqs ’
s
which is required.
306 7. Lagrange’s differential equations

Considering a function of the generalised coordinates,


generalised velocities and time
we introduce the notation

(/) = df
yj>
dt dqs _ (s = 1, . . . ,n), (7.1.8)
dqs
which is referred to as Euler’s operator over /. In particular, if /
does not depend on the generalised velocities, then
df_
M/) = - dqs' (7.1.9)

It is clear that
£s (fi + /2) = £s (/i) + £s (/2) •

(7.1.10)
Using Euler’s operators we cast Lagrange’s equations (4) in
the form
ES(T) = QS (s = l,... ,n),
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 4>, Lagrange’s equations, by virtue of
eqs. (11), (12) and (5.11.12), take the form
d$
£S(L)=QS~ — (s = 1, . . . , n).
In the case of non-holonomic constraints (5) Lagrange’s
equations is written as follows
i d<
£ (-T') — Qs 4“ ^ ^ (7.1.14)
s
S> {s = 1, . . . , n ) .
Afc&fc dqs
g k=1
The terms
i
^A kaks

(7.1.15)
k=1
in eq. (14) present the generalised constraint forces of the
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 coordinates, 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
<is = ,q n ',qi,- -- ,qn;t). (7.2.1)
An explicit form of these relationships is derived below, cf. eq.
(7.4.7).
ip(qi,--- ,qn;qi,-- - ,qn;t). (7.2.2)
Its time-derivative constructed by means of the equations of
motion is given
by
dip _ dip ^ f dip . dip ..\
dt dt + ^ \dqs (is + dqs / ’
in which the generalised accelerations are replaced by
quantities (1) from the equations of motion, i.e.
dip dip f dip . dip
(7.2.3)
~dt = ~dt +
s= 1 dqs dqs
^ 1 A^~-qs +
This function ip 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 ip 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
dL _dL ^ fdL .. dL \ dL ^ / dL .. d dL \
dt dt \<9gs qs + dqs / dt + “ Qs + Qs dt dqs )

^ ^ QSQS ^ ^ ^k ^ ^ ttksQs 4“ ^ ^ Q . 4s-

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 Tv L Tv
^ ^ QSQS 4“ ^ ^ 4“ ^ ^ •
dt Qs
6=1 k=l 6=1
308 7. Lagrange’s differential equations

or, by virtue of eqs. (6.4.19) and (5.10.3),

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 . Then the right hand side of equality (4) is
zero and we arrive at the following equality
(7.2.5)

where h is a constant value. The assumption that L does not


contain t explicitly and then all = 0 is not equivalent to the
assumption of stationary 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 expressions for ps and L. Instead of eq. (5) we have
2T2 + T\ — (T2 -J- T\ -}- To) + n = ft or T2 + II — T0 = 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 + n = 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,
(7.2.8)

where N* denotes the power of the non-potential forces


except for those which are expressed by the dissipation
function and
(7.3.5)
<dBk dBs\ .
dqk ) qk'
\ dqs

The quantities
s
- dBk _ dBs dqs (7.3.6)
T k. dqk ’
0 7l2 ‘' * 7ln
721 • 72 n
7 = 0
Tnl 7n2 * ' ■0

are named the gyroscopic coefficients whereas


n
rs = $>sfc<?fc

(7.3.8)
k=1
denote the generalised gyroscopic force and 7skffi as the gyroscopic forces.
The equations of motion (4) are now cast in the form
£s(T2) = r s + Q s - — --5f (s = l,...,n), (7.3.9)

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
Q s . In this case the generalised velocity qs does not appear in
the s — th equation, as = 0, and the coefficients of 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 nn
r
N' = E ^ = EE 7 skQkQa = 0.

7.3 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

- (14:-4;)
d A• 1 dAkm
.
~QkQm
d q s
k =1

A •• , 1 dAkm \ . . dAks .
k =1 L+XX( - 5ITJ k—1
+
X
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
1 0Akrn
2 dqs
4k4m
dAkm nn
1 dqs qkQm = ^2 [*> S] 9fc9m,
2 k—1 m= 1

where
dAks dA dAkm \
[k, m;s\ = - ms
dqm dqk dqs )
denotes Christoffels’s symbols of the first kind (introduced in
Sec. 4.10) for matrix A of the coefficients of the quadratic
formT.
The equations of motion (3.9) can be recast in the form

k=l m= 1k—1

rs + Q s an* dB„C'-Ds
(
^ dAks. \ ^
qk (s= 1 , . . . ,n).(7.4.1)
dqs dq dt
k=1
As mentioned in Sec. 7.2 these equations present a linear
form in the generalised 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 introduced in Sec.
4.2. Their elements are (see (A.3.13))
A
f = ^ (l,s = l , . . . , n ) , (7.4.2)

where Ais denotes the algebraic adjunct of the element Ais of


the determinant | A| of matrix A .
Multiplying each equation in eq. (1) by A ~ j and summing up 1

these products over s we obtain


n nn
qi + X>s‘££ [k, TO; s] qkqm
s—1 m=l k=1
= E-4;1 rs + Q s
an dBs \ dAks .
(7.4.3)
*
dqs
since
53 ^ks^si 1 xA
= f>ki
f0 k 7^/,
s=1 ^ 5 v \1 k = l.
312 7. Lagrange’s differential equations

Introducing the values


= = (74
'4)
referred to as Christoff el’s symbols of second kind (also named the
braces) for matrix A, we arrive at the equations of motion in the
generalised coordinates resolved in the generalised
accelerations

4 +EE l (7.4.5)
k QkQm
k=1 m=1 m

Z^Asi ( s + <2s hn
r
2 ^ Qt ) (l l , . . . , n ) .
9<
,=i \ 1> k=X

In the case of stationary constraints Lagrange’s equations


(1) and (5) simplify and take the form
n nn
'^2AksQk + '^2'^2[k,m-,s\qkqm = Qs - — (s = l , . . . , n ) , (7.4.6)
k= 1 k=l m=l a s
^

k=1 m= 1 v J s=1 v
^

Qs + (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' = y/mx, y! = y/rriy, zf = yfmz,

(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 y/m, such that
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

H»2=KS)2’ --
(7 5 3)

where ds is the” arc element” of the particle trajectory.


Introducing the curvilinear coordinates g1,#2,#3
r = r ( g \ g 2 , < 73 ) (7.5.4)
we obtain
r = v = rsqs, (7.5.5)
rs denoting the basis vectors, see (B.4.4). The redundant
summation sign is omitted due to the convention of
summation over the repeated index 5. 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
T = |rs • rkqsqk = \gskqsqk. (7.5.6)
The coefficients of the kinetic energy (earlier denoted as Ask)
play the role of the covariant components of the metric tensor.
This enables us to write
{dsf = 2T{dtf =gskqsqk. (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
<5'W = F- -$= = -^F - r s 6qs = Qs6q
s
.

According to the above definitions Qs are the generalised


force and the covariant components of force Q. The
contravariant components are given
by
Q s = g sk Qk- (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
r +
w
=*=+=r* ( •
314 7. Lagrange’s differential equations

The quantities in parentheses are the contravariant


components ws of vector w
ws = qs + (7.5.12)

Under the introduced notation for vectors w and 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
qs + = Qs = gskQk- (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 conventional form (1.11) are obtained by
using relationships (13) in terms of the covariant components.
Indeed,
ws = gskWk = gsk (^qk + j^”1^ = gskQk + [m, t; s] qmqt. (7.5.15)

Taking into account that

gSkqk &9sm .' &


dgmt
[m, t] s] <fV qmqt
dt 9smq
F<1 9
~dq s 2 dq s
m
and using expression (6) for the kinetic energy we obtain
d dT dT
(7.5.16)
dt dqs dqs'
Thus, Newton’s second law in the covariant components
takes the form of Lagrange’s equations
d dT _
(7.5.17)
dT_ dt dqs
dqs
or, in the explicit form (4.6),
gskqk + [rn, t] s] q171^ = Qs. (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
(7.5.19)
■k <kk.
V = Tkq = r k — s,
7.5 Geometric interpretation of particle motion 315

where ds is the arc elements defined by eq. (7). The vector


dnk
T = r
k~T~

(7.5.20)
ds
having the contravariant components dqk/ds is directed along the
trajectory tangent. It follows
. from
. eq.
.. .o (7)
dr
that r is the unit
V = TS, w = V = TS T S ——. (7.5.21)
ds
Now we have
dr d2qk dqk dr& dqm / d2q f k 1 dq dq \
k m f

ds k ds2 ds dqm ds k \ ds2 \mt) ds ds ) **


(7.5.22)

This vector is referred to as the curvature vector. The quantities


fcr _ fr\ dqrn dqf
ds2 \mtj ds ds
are its contravariant components whereas the quantity
k = V9smkskm (7.5.23)

is the curvature of the trajectory. The unit vector n represents


its main normal. Hence, we obtain
dqk .. ( d2q r
k \ dqrn dqf \ .
w = TS + s
2
kn = Tk + (7.5.24)
ds V ds2
S
^mtj ds ds
The acceleration vector is represented as a sum of two
components, namely the tangential W(r) and normal W(n)
accelerations
dqk .. S = (d2qk \k\dqrndqt+x. 2
W =r
« ‘rf7 ' (7.5.25)

Decomposing vector Q in a similar way


Q = Q(r) + Q(n)> (7.5.26)
we obtain by eq. (20)

dq dql dq* dqk


Q W = r Q - r = rk^-Qsrs ■ rt— = Qt — —rk (7.5.27)
ds ds ds
316 7. Lagrange’s differential equations

and furthermore
rkQt ( gkt dql dqk \ ds
Q(n) = Q - Q(T) = (V
ds ) ' (7.5.28)

The natural equations of motion are thus written in the form


dq* rfV f t \ dqm dqk ds2 dql dqk\
— Qt
_ \raA; / ds ds ds dsj
ds ’ (7.5.29)
In the case of the potential forces the first term can be
transformed as follows
.._ds_. _ d_s^_ _ _dU dxf_ _
(M ds ds 2 dqf ds ds
Integration yields
s2 = 2 (h - n), (7.5.30)
where the integration constant h is the total energy of the
particle. Eliminating s2 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

dV f t } dqmdqk _ 1 cM ( kt dq* dqk \


ds2 \mk j ds ds 2 (h — II) dqk \ ds ds J ’ ^’‘

Of course, they must be considered along with the following relationships


dql dqk 1
9kt (7.5.32)
ds ~ds ~ '

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))
p = p(q1,q2), (7.6.1)
where q denote the Gaussian coordinates. In what follows,
a

the Greek
indices take values of 1 and 2. The (7.6.2)
velocity
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

V = P a (qa + | ^ | + mba0qaq0, (7.6.3)

where m denotes the unit vector of the normal to the surface


and bap 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 components aap are the
coefficients of the first quadratic form of the surface
ds2 = CLapdq^dqP = pa • p/3dqadq/3 — 2Tdt2, (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

wa = q a + (7.6.5)

obtained by covariant differentiation of the contravariant


components of the velocity. The second one is the
acceleration normal to the surface
m , .a.n .2 , dq dqP ,2f
w (m) oaaq q = s mbaa—--------— = s km. (7.6.6)
p
ds ds

Here, due to (B.8.4), the value of k denotes the normal


curvature
k = bCL
apdq dq
a 0 (7.6.7)
af3dq dqP
a

The covariant components of the accelerations on the


surface can be rearranged into the form

wa = aa/3W0 = aapqp + [/?, 7; a} (fq1 ddT adT a


dtdq dq ' ^''

Introducing the independent variable s in expressions (5) for


the contravariant components wa of the acceleration vector w on
the surface we obtain
„dq a
d2qa ( a 1 dqP dq1
w W Pa Pa
=
S
~d7 + ds2 \/^7j ds ds
Pa I d + S2k*a (7.6.9)
qc
d
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* determines the direction of the
geodesic normal of the trajectory. Now we have
w =ST + S £;*IT,
2
(7.6.10)
where the geodesic curvature £;* is given by eq. (B. 8.8). The
vector of acceleration on the surface is the sum of two
vectors, namely the tangent acceleration 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
S'W = —L (F + R) • <5r = -LF • pa6qa = Qa6qa.

(7.6.11)
\/m yjra
As before, we introduce the vector Q with the covariant
F = F pa+F(m)xn.
a
(7.6.12)

Newton’s second law now takes the form


V =-L (F + R).

First, we arrive at the equations of motion

Wa = q a + jj=a a0
Q/3 (7.6.13)

or, in the covariant components,


d dT dT
(7.6.14)
dt dq° dqa ‘
Second, we obtain the expression for the normal reaction force

R = ( ^(m) + m, (7.6.15)

the presence of the factor yjm 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 equations of motion in the form corresponding to the
natural equations of motion in projections on the tangent and
geodesic normal of the trajectory
dqa
Qa
d 2c ' a 1 dqP dq7 \ 2
ds ’
dqP (7.6.16)
+ = Qp
dqa ds
ds2 /?7 / ds ds J
If Q = 0, the velocity remains constant.dsWe 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 form3 3
1
x + 2 3 1 2
{aa}* * = % l — 2rx x — x r (x ) ,
3 2

w — x

w x
2
+ 2rx l x 3 2 2
— X T (X3) ,

w= x3,
where r 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
du
Qi =0, Q‘z = 0, Qs = -7^3 = -(]■
The contravariant components are equal to
1
Q = -gg 13 = -grx2, Q2 =-gg23 =-grx\ Q3 = -gg33 = -g.
320 7. Lagrange’s differential equations

The equations of motion take the form

x1 — 2rx2x3 — xV2 (x3)2 = —grx2 '


x2 + 2rx1x3 — x2r2 (x3)2 = grx , ►
1
(7.7.1)
x3 = ,

In order to write down the equations of motion in terms of the


covariant components we construct the expression for the
kinetic energy
T \s2 = ^gikx'x* = 1 {(i1)2 + (£2)2
2TX2X1X3 + 2TX1X2X3 + |^1 + r2 (x}2

Then we obtain

x1 — rx
2 3
x — 2rx2x3 — T 2 X ( X 3 ) 2 = 0, x2 +
X

rxxx3 + 2rx x — x2r2 (x3)2 = 0,


1 3
(7.7.2)
d ^£l2 + X22^
2
(xxx2 - XXX2) |

dt {[
l+T
= -9-

It is easy to prove that the projections of acceleration w of


the particle on the rotating axes Ox1 and Ox2 are equal to the
covariant components W\ and 1^2, 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 VQ of the initial velocity lies) coincides with
plane Ozx and the origin of the coordinate system is placed in
the initial position of the particle, then we have
1 2
2 = v0zt - -gt
X —

and furthermore

x1 = VQXt cos TX
3
, x2 = — foa^sinrx3, x3 = VQ zt — - gt2.

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 q1 referring to a
orthogonal curvilinear coordinate system is given by

dhl
. o..i T dh\ /.i\ 2 dh\ .i .o 0 h a 1n3
M +hl
w^ ^ q +2
W hiq q
V - i? q

h2
|^r (<72)2 - («3)2 = (7-7-3)

where hi denote Lame’s coefficients. The other two equations are obtained by an
appropriate change of the indices. Expressions (B.10.8) for Christof- fel’s symbols are
used here. Equations in terms of the contravariant components, i.e. in the form of eq.
(5.14) are obtained by dividing eq. (3) by coefficient h2.
As an example let us consider an orthogonal system of curvilinear coor dinates q1
= 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 ( l ) = f ( a + i(3), (7.7.4)

where / (7) is a function of the complex variable 7 = a + i/3. Its derivative f (7) is
assumed to be non-zero in the domain of the complex variable 7. Then

dx + idy = f (7) (da + id/3), ds\ = |/' (7)|2 (da2 + d(32) .

Therefore, Lame’s coefficients for the isometric coordinates ha and hp coincide and are
equal to

hl = hl = I/' (7) I2 = f (7)/ (7) = h2 (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

(■ds)2 = h2 ( a , P ) (da2 + dp2) + h2v (a, p) dip2,

where denotes the azimuthal angle and

^ = |Im/(7)|2. (7.7.6)
322 7. Lagrange’s differential equations

The equations of motion (3) in terms of the coordinates


a, /?, <p have the form
dlnh (*-*■) d In h • h(n dhm . o Qi
a+ +2 /12’
da d(3
9 In h V - * ) dlnh . •ah0 v dhvv .2= Q2 (7.7.7)
(3 +~w +2 ~a^ -»W v’
Jth>= °'

where it is assumed that Q^ = 0.


The latter equation yields the integral of
area
(7.7.8)
hl<p = kv,
where k^ is an integration constant. Removing (p by means of
this relationship and assuming the existence of potential forces
we obtain
dlnh (. o ^2\ dlnh . •
(a - P J + 2 Q ( 3 a(3 -
Id/ k2
— (II +
a+ da d(3 h2 da 2 h%
9 In (7.7.9)
h V-*) dlnh
da h d(5 ( 2h
a(i 2 11 + 2

/? + +2

Now, using the energy integral


7i2 (&2 + |32) = 2 (fio - n - i) , (7.7.10)

it is easy to rearrange the first equation in (9) into


the form
S + a(1n/,T = i [ £ ( E 0-n-||'
+ din h Eo-IL-
2

da 2/.2
The second equation can be written by
analogy. Hence we obtain the system
of equations
It„V2
dt (k2°)=^h
H'l
(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 (3 separately. This is, for instance, the
case for planar motion of a particle attracted by
7.7 Examples 323

two centres. Placing these centres F\ 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(3) = c (cosh a cos (3 + i sinh a sin /3).

Then
2
h = c2 (cosh2 a — cos2 (3) .
The distances n and 7*2 from particle M to the attracting
centres are found using the equalities

r\ = c |cosh (a + i/3) + 1| = c (cosh a + cos (3),


7*2 = c |cosh (a + i/3) — 1| = c (cosh a — cos (3).
Taking into account that the potential energy is given by
fl h 1 cosh a - (/1 - /2) cos/3].
n= r\ T2 [(fi + ff)
cosh 2
a— cos2 (3
where fi and are constant values and &+ — 0 for the plane
motion we obtain
TT\ _ 2 I IP______u2„ , /1 + / 2
h 2 (E Q — II) = c2 E0 cosh
c 2
a 2
c2 ( E Q COS (3 + cosh a I —
c
h-h cos (3

Integration of equations (11) yields


hYa2 = 2c2 ( EQ cosh2 a + ^ ^2 cosh a + ) ,

h4pA = —2c2 ( E Q COS2 (3 + —-------—


cos/3+ 72 .

Adding these equalities and using the energy integral is it easy


to find that 7i + 72 = 0. This we arrive at the system of two
equations
(da) (d(3Y
EQ cosh a + Z1 cosh a —
2
—#0 cos2 (3 — ——— cos (3 + 72

2 (dt)2
(cosh a — cos'! / ? )2
2 ’
324 7. Lagrange’s differential equations

which yields another two integrals

>

(7.7.12)
The solution of the problem having four integration
constants to, C, Eo, 72 is thus reduced to calculating the above
elliptic integrals. In passing we notice 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
h2 — h\ (a) + (P), h2II (a, (3) — (a) + ^2 (P) (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
(.E0h\ - J
- 7) 1/2 dot- (Eoh% - *2 + 7) 1/2 d(3 = C, '

Jh22(E0^2-*2+ 7) 1/2 d(3.


(7.7.14)

7.7.3 Equations of motion on the surface of revolution


These equations have the form

(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
Q° = -9 J - S I Q v = 0-
7.7 Examples 325

Integrating the second equation and eliminating Cp2 from the


first equation we obtain
r <p = ro¥>0) s + = -5^- (7-
7.16)
The latter equation yields the energy integral
s2 = 2^h-gz-^y (7.7.17)

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 quadratures. In the case of a
spherical pendulum
2 = Rcost?, r = Rsin d = yjR? — z1,
and the obtained differential equation
i2 = 2 (/i — gz) {R*-z*)-rtti
can be integrated in terms of the elliptic functions. The motion
of a particle on the surface of a circular cone
2 = s cos a, r = s sin a,
the surface of a paraboloid of revolution
r2 = az
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 surface. 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 g 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
326 7. Lagrange’s differential equations

where the negative sign (force of attraction) corresponds to


the lower position of the cone vertex and the positive sign
(force of repulsion) to the upper position of the vertex. The
angles ^ and tp (the azimuth on the cone surface) are related
by ^ = (psinai. The result is

s2 = 2 (h qp gs cos a) — 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 rrii by quantities proportional to
the Cartesian coordinates of the fixed basis system
Csi—2 = VmXi, £3i-1 = y/rniyi, Gi = y/mzi, = ,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%N. The position vector of this representative point is
denoted by r, and quantities are understood as its projections
on axes of the Cartesian coordinate system in E3N. The
relationship
r = r(q\... ,qn) (7.8.2)
expressing the Cartesian coordinates in terms of the
generalised coordinates g1,...,#71 defines a Riemannian manifold
Rn in E3N. 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 dr of Rn is determined by the
(7.8.3)
dq“'
where the Greek indices take values 1,..., n. The vector
represented as aara belongs to Rn and aa denote the
contravariant components of this vector. Now, differentiating
eq. (2) with respect to time we obtain the velocity vector
r = v = raqcx (7.8.4)
7.8 Geometrical interpretation of the equations of motion of the system 327

belonging to Rn. Here qa denote its contravariant


components. The kinetic energy of the system of
particles is given by
1 N sS> 1 3N ,
i=1 (il+Vi+zf) = r=-V
U=1
Expressing v by means of formula (4) we obtain
T
= \r<* • r0QaQ13 = \aa0qaq0, (7.8.5)

where aap denote the contravariant components of the metric


tensor of space Rn.
The active forces F* 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 = Qa Ta,
ra 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 F* due to virtual displacement of the system particles is
equal to the elementary work of ” force” Q due to virtual
displacement <5r
Q 6r = Qar° ■ vfj6q0 = Qaa$6q0 = Qa6q .
a

The acceleration vector is defined as the following vector


w = wara — wara
belonging to Rn and having the contravariant components wa.
They are obtained by covariant differentiation of the
contravariant components qa of the velocity vector v with
respect to time
^ + {/fryP®7- -'
(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 wa are calculated due to the
known rule
Wa = aaf3w0 = aa/sq0 + [f3,7; a\ q0q1. (7.8.9)

Repeating the calculation of Sec. 7.5 we can write this


expression in the form
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

qa + {= aa0Q0.

(7.8.12)

The result is the equations of motion in the form (4.7)


resolved in the generalised 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-
s-Q d2qa ( a 1 dqP dq7 a (3 dqa
as ds2 ^ \/3^) ds ds dqP I Q(3-

ds ds (7.8.1
3)
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 2 a
d q f a 1 dqP dq7
(7.8.14)
ds2 ^ \/3^j 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 Q(i), Q(2)> •••, Q(m) acting
separately. The kinetic energies of each of these separate
motions are denoted as T^), T(2),..., T(m). Then the same
trajectory is realised when the above forces Q(i), Q(2), •••, Q(m)
ac
t 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
dT(i)
ds T + 2T{i)kn = Q{i) (i = 1,... ,m), (7.8.15)

where r 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 M\ and M2. The
lengths of the inextensible strings OM\ and MIM2 are l\ and /2,
respectively. The suspension point O moves in the vertical
plane OMIM2. The velocity of point O is denoted by Vo and the
angle between Vo and the downward vertical is denoted by a.
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
T = ^ [(mi + m2) l\Cp\ + m2/2(,b2 + 2m2/i/2(Jb1(Jb2 cos (cp1 - <p2)] +
^ l
Vo [(mi + m2) l\Cpx sin (a - ipx) + m2/2(,b2 sin (a - tp2)] + - (mi + m2)
VQ
= T2+TI+T0.
(7.9.1
)
£1 (T2) = (mi + m2) /fv>i+m2/i/2V,2 cos 0Pi ~ sin ((p1 - <p2)
£2 (T2) = m2l%ip2 + rnvhhip-i cos ((p1 - tp2) - m2lil2(p\ sin ((p1 - <p2) ■

While calculating £s (Ti) the quantities Vo and a are


considered as pre-
scribed functions of time. The result is
£1 (Ti) = (mi + m2) h [*o sin (a - ipx) + v0dcos (a - y?i)],
£2 (T\) = m2/2 [v0 sin (a - ip2) + ^o^cos (a - tp2)] .

The gyroscopic terms do not appear since the coefficient of


<p1depends
only on and that of Cp2 only on ^2- The term T0 does not depend
on the generalised coordinates and the corresponding Euler
operators are
II = —m\gl\ cos tp1 — m2g (l\ cos + /2 cos ip2), (7.9.3)
330 7. Lagrange’s differential equations

and the equations of motion have the form


(mi + m2) Zi<£i + m2ZiZ2<£2 cos O^i _ ¥2) + '
m2l\h^p\ sin (cp1 - (p2) = - (mi + m2) h [vo sin (a - cpi) + u0dcos (a -
ipi)} - g(mili + m2Z2) sin +Qi, >
m2Z2(£2 + rri2lil2<Pi cos (^ - (p2) — m2/i/2^ sin (ip1 - (p2)
= —m2Z2 [i>0 sin (a - <p2) + uo^cos (a - <p2)] - #m2Z2 sin (^2 +
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 Mi and M2 from the position of relative equilibrium.
Then, setting in eq. (4)
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 (7.9.5)
Ull^l T ^12^2 H" bn&i + ^12^2 T
where
1 + A IA2
an = ( m i + m 2 ) Z i ,a i 2 = a 2 i = m 2 Z i / 2 , a22 = m2Z2,
\J{1 + ^i) (l + A2)
2 + A IA2
2 _|_ A
bn = (hi + &2)^oZi-—;—bi2 = b1 +2i=k
A 2
^oZiZ2,
"“ “ Y(1+Ai) (1+A")
z>22 — &2^0Z2-- -f, c2 77125/2 ^/i + A2,
1 + Ao
CI = [(miZi+m2Z2)^ +(mi+m2)^ZiAi]
A+A1
Here — Ai and — A2 denote the tangents of the angles of the
strings in the position of relative equilibrium
ki + /c2
(m i + m 2 )g “ Ql
Ai = U A2
m2g
Assuming the absolute values of these angles are smaller
than 90° we have
7.9 Example of applications of Lagrange’s equations
331

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
=
2 ^a12^1^2 + ^22^2) , n* = - (ciE± + £2^2) 5
= 2 (&1^1 + 2&12^1^2 + ^>22^2) •
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
jt{T* + n*) = -2$*,
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 £, 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,
( \
1 i
.. 1
U=
2 . 2m l+u
l H------u
\P /

It has a particular solution u — 0 corresponding to the case in


which both end of the string have initial velocities xo = Vo- It
follows from eqs. (6) and (8) that x — y — —g. The string moves
vertically with a constant acceleration and the length of each
part does not change. This trivial solution is of no interest.
Let us return to eq. (9) and take ii as an independent
variable,
.. du .thus
u = —u.
du
7.9 Example of applications of Lagrange’s equations 333

As u 7^ 0, eq. (9) admits a separation of variables


du 1 du 1 du
u 2l+u +
2 _ 2m
Z + —
P
Let us assume that the length of each part of the string is
1/2at t = 0, while the difference between the initial velocities of
the right and left parts is positive, i.e.

t — 0, u ~ 0, u = UQ > 0.

Integration yields

l
(7.9.10)

It is clear that the value of u cannot exceed l. Moreover, u


increases mono- tonically from zero and reaches l at the time
instant when the string straightens itself.
An expression for t in terms of u is easily obtained by
integrating eq.
(10)

u0t = 1 - du. (7.9.11)


2m
l +
PJ

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 s[ < £ 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’[g — mg or S’[ = {ps" + m) (x + g).


334 7. Lagrange’s differential equations

Replacing x + g by means of (8) and considering relationships


(9) and (10), we obtain

o/1_ _ Ei L,
* 21 2 ’
/ 2m
(/ + u) I / H-----u
(7.9.12)
ii / 2m \

^ = 2 (Ps i + m) 2 “0*
(/ + u) (/ + —u

For a section of length 52 of the left part of the


string
27 (1
m2
^ (11 + u)/ f (/l +
H--- — «
S 2 = pS 2 (y + g) = PS 2 (x + g-u) = -puls 2 P
(7.9.13)

By virtue of (7) one obtains that at point K


s" = z - X = 1 (l - u), S2 = Z - V = 1 (/ + w)

and
/ 2m
Mz + —
_ c __ 1 ■ 2___V____
c" _52
5l
“ “ 4^° , /. 2m (7.9.14)
(/ -|- U-) | / ~b ---- — U

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
As shown in a recent paper by W. Steiner and H. Troger
1

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
a massless U-tubefold
by in the case
means of abalance
of the zero radius of the fold.
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

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 intersection of the suspension axes. The kinetic
energy of mass m is thus
(7.9.15)
which implies that the only correction is to add to the
mh2

moments of inertia A2 and B2 of the gimbals. We denote A'2 = A2


+ , B2 = B2 +
mh2 mh2.

The position of the axes is shown in Fig. 7.3. Axes 0£r]


( denote axes attached to the fixed base of the facility, axis
being the axis of rotation of the outer ring. The trihedron of
unit vectors is related to
the
inner ring. In the case a) axis is horizontal whereas in case b)
II = mg( = mgh cos a cos [5,
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

(ai + b cos2 (3) a — ba/3 sin 2(3 + C3 (cp + d sin (3) '(3 cos
(3+ "j
C3 sin(3 ((p + dsin/?)* = - —, (7.9.16)
aa
^2/? + -&d 2
— C3 (<P + dsin/J) acos/3 =
— —,
C3 (<p + dsin/3)* = 0, ,
where
CL\ — A\ + C2, U2 — B2 + A3, b = A-2 + A3 — (72-
The latter equation admits the integral (7.9.17)
C3 (</? +ofdsin/?)
expressing the conservation = C3r0,momentum of the
the kinetic
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
<911
+ b cos2 0) a — bd(3 sin 2(3 C T P (3 - ,
(ai + S Q = —
da ’
COS

(7.9.18)
1 2 on
a2(3 + -6d sin 2(3 — Csroacos/3 — — — .

The terms — C^TQ(3 cos (3 and C^r^acos (3 play the part of the
generalised gyroscopic forces. Though the constraints are
stationary and term T\ 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

r+n (ai + b cos2 /?) a2 + a2(3 + U = E0 (7.9.19)


the term being included in EQ.
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
dt da
= 0,
yielding the following integral
dT (a\ + b cos2 /?) & + Csro sin (3 = .
da (7.9.20)

Here K% denotes projection of the resultant angular


momentum K° on axis O£. 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 [3 are small and the linearised
(ai + b) a + Csro(3 — mgha, a2$ — C^r^a = mgh(3. (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 = D\ cos (kt + xjj), (3 = D2 sin (kt + ij)).
We obtain two linear homogeneous equations in D\ and D2 [ ( a i
+ b) k2 + mgh] D\ — CsrokD2 = 0, C^r^kDi — (a2k2 + mgh) D2 = 0. The
nontrivial solution exists when the determinant of the system
is zero,
A (A;2) = [(ai + b) k2 + mgh] (a2k2 + mgh) — C^r^k2.

(7.9.23)
If h < 0 which means that the mass m is on the underside, then

A(0)>0, a
A f - ^ < 0 , A (co) > 0,
i + bJ
\ \ a2 J
and equation (23) has two roots. Let k\ < k2, then A;2 is less
d\ + b
that the
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

[Cf7*0 — (ai + b + <12) mgh]2 > 4m2g2h2 (ai + 6) a2-


The expression in brackets should be positive otherwise
both roots k\ and k\ are negative. We obtain the following
inequality
C3 \ro\ > mgh (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 the general solution of the system of
differential equations (21) is given by
a D[^ COS (k\t + ^1) + D^ cos (k2t + ^>2) 5
P
(1) (eg + b) k\ + mgh
D\ ■ sin (kit + ^x) +
C^r^ki
^(2) (ai + b) k\ + mgh . . ,

The four integration constants D[X\ D^\/ip1,/ip2 are determined by


means of the initial conditions. The motion is a superposition
of harmonic oscillations with frequencies k\ and 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/? = w,
2• (1-n2) 2 (EQ — mghu) —(Kz - Csr0uf2 a\ + (7.9.25)
b (1 — u ) ’
2a u
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 ^ 0. Analysis of the
stability of
the vertical position /? = ir/2 of the rotor, see Fig. 7.2b, is given
by Magnus
[61] and Rumyantsev [78]. (7.9.26)
In the case of the vertical rotation axis of the outer ring, the
7.9 Example of applications of Lagrange’s equations 339

The latter equation along with eq. (18) yields a particular


solution
/^ = 0, d =
describing a uniform rotation of the outer ring and the vertical
position of the rotor axis. Taking

and linearising eq. (26) under the assumption that e is small


we obtain a^e + ( — b a o + Csr0ao — mgti) e — 0.
The motion of the rotor axis will be a harmonic oscillation
about the vertical position under the following condition
/ (d0) = ~bao + C3r0a0 - mgh > 0.

(7.9.27)
The quadratic equation / (do) = 0 has real-valued roots (do)i
and (do)2 if
Cf 7Q (do)
> 4bmgh < =
x
4do
mgh (A!<
2 + (do)
As —
2
C2)•

(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 (fi and ^ between the rods A\B\ and
B\C\, respectively, and axis A\x are taken as the generalised
coordinates. The difference — (p1 is denoted as f31. The
moments of inertia of the rods A\Bi and B\C\ about axes
perpendicular to plane A\xy and passing through point A\ and B\
are denoted by Ji and 0i, respectively. The lengths of the rods
A1B1 and B\C\ are denoted as v\ and p1? respectively, and the
distance B\G\ between the centre of gravity of the second rod
and joint B\ is designated as S\. The force Fi applied at point Bi is
perpendicular to AiB\. The angular accelerations of the rods
T= \ (mirlfti + 2m1r1s1(p1'ip1 cos(31 + ©i^i) . (7.9.30)
340 7. Lagrange’s differential equations

FIGURE 7.4.

The first term is the kinetic energy of rod A\B\ rotating about
the fixed axis A. The other terms represent the kinetic energy
of rod B\C\ calculated by means of eq. (4.7.7).
The elementary work of the force Fi is F\r\bipx, that is

Qi = Fin, 0.
Using the notation
An = J\ + m\r\, A12 = m1r1s1 cos/31, A2 2 = &i (7.9.31)
we arrive at the equations of motion
£<?! (T) = Mwi + ^12^1 + ^7^1 = Fin,
( T) = A12&1 + A2 + A = 0-
Olf1 (7.9.32)

For the system under consideration, equations in (4.6) have the form

An^ + Midi + [1,1; 1] ¥>i + 2 [1, 2; 1] + [2,2; 1] fa = Qu j


A2m + A22^1 + [1,1; 2] $ + 2 [2,1; 2]
+ [2,2; 2] = Q2, J
(7.9.3
3)
indices 1 and 2 being referred to and Comparing eqs. (32)
and (33) we see that the non-zero brackets are
[2,2; 1] =dA12= —TOiriSi sin/Jj, [1,1; 2] = mirisi sm/Jj. (7.9.34)
dip i
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

*’■ = A l A22Fin + mirisi sin P1 ^1^-12 +'01^22^


(7.9.35)
Ai2Fin + 77iirisi sin J

/?1 (&\Ml
where A = An A22 — A\2- 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 elimination of constraint multipliers


Let us return to Lagrange’s equations (1.6)
A

£s (T) = Qs + Akaks (s = 1,... , 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

J2aksQs= 0 (k = l , . . . , 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,... ,77)


d&kr
are satisfied, then the k — th constraint equation can be
replaced by a finite equation
F k ( q i , - - - ,qn) =0, (7.10.3)
where
dFk
&ks —dq ' (7.10.4)
s

The number of redundant coordinates is equal to the number


of integrable equations in (2).
anai2 .. • din
0* a 22 • • • d2 n
21
an an • • • din
dn di2 • • • dn
d2l d22 • • • di2
n* —

dll d2l • • • du
(7.10.5)

is equal to zero.
Then one can determine the constraint multipliers from l
equations, for instance the following ones
i
J2Xkaks = £ s ( T ) - Q s ( s = l , . . . , l ) . (7.10.6)
k =1
It follows from this that the square l x l matrix

(7.10.7)

is non-degenerating because it is the transpose of the l x l


submatrix a* of matrix (5). Let a q r denote the algebraic adjunct
of element a q r of r — t h row and q — t h column of the
determinant |a*| of matrix (7). Then
1^
rr|Eafcs l £ s ( T ) - Q s ] , (7.10.8)
\a*\ s= 1
and substituting these values of A k into the other (n — l)
equations in (1) we find
i
£l+r (T) = Ql+r + Y, [£, (T) ~ Qs\ Nt+r (r = 1,... ,n — l), (7.10.9)
S= 1

where
N
1 ^
l+r = T-nYakSa^+r- (7-10.10)
I S=1

The latter sum is the determinant obtained from |a*| by


replacing the s — th row with the following row

&l,Z+r i • • • i &l,l+r• (7.10.11)


7.11 Examples 343

The sum
l
Q*S = Y1 (« = !»•••,«)

(7.10.12)
k =1
is the generalised constraint force corresponding to coordinate qs. In
order to understand the reasoning for this name we mentally
eliminate the constraints described by eq. (2) and include
their reaction forces in the active forces. Then the s — th
equation in (1) can be cast in the form
n
^ ^ tiks^Qs — (7.10.14)
0)
s=1
and, by virtue of the theorem of Sec. 6.1, the quantities Ql are
represented by formulae (12).
We conclude now that the generalised constraint forces are
defined by the following equalities
Q*s = Es (T) - Qs (s = 1,... ,/),
(r = 1,... , n - l ) . (7.10.15)
Qhr=ENUr[£s(T)-Qs]

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 cp1 and <p2. We also introduce two
redundant coordinates, namely angles i/q and which are
related to yq and cp2 by the following relationships
= 7*1 COS (p1 + px COS Vq — T2 COS (p2 — pr 2 COS (7.11.1)
l\)2 = 0,
$2 = T\ sin yq + px sin i/q — r2 sin ip2 — p2 sin ip2 — a = 0,
obtained by projecting the pentagon A1B1CB2A2A1 on the
coordinate axes.
344 7. Lagrange’s differential equations

FIGURE 7.5.

The kinetic energy of the system due to eqs. (9.30) and


(9.31) is given by

T=— + 2A[<2 +^22^1) “b

\ (^11 ‘fit + 2^12 ^2^2 + A


22^l)

(7.11.2)
where the subscripts (1) and (2) are respectively referred to
the rod pairs AiBi,BiC andAs F1A.A1B1 and F2_LA2F?2, the
generalised
forces are equal to

QVx = Pin, QV2 = F2r2, 1= 2 =


°-
(7.11.3)
£<Px (T) - QVl = + ^12^1 - minsi^i sin
A^iVi
- '
£ij>x (T) - Q^x = ipi + A^ip1 + mirisivl sin/Jj, ^
£<P2 (T )- QV2 = A$ip2 + ^i2 ^2 - m2r2S2^2 sin/?2 - £^2 (T) - Qv,2 =
)

(
A$(p2 + A $i>2 + sin /?2, >
(7.11.
4)
Adopting the following numbering of coordinates
91=^1, Q2 = ^2 93 = ^1, 94 = </?2> we
can cast matrix (10.5) in the form
_ — pjsin^ p2sin'ip
r cos
2 — rising r2sin(p2
/^COST/q — p 2 C O S 2 p 2 l (f\ —r2 COS (p2
= 1 —rx sinc/q 7*1 cos
ipx
Kl p2 sin tp2 —P2 cos
1 -p sin^q ip
x px2 cos 2px
NP =
Kl —rx sin ip rx cos ipx
x

lalogy
sin (V>2 - P 1P 2 sin (V’I

c7
_ r2

II
< p 2 ) P i sin

= n sin (y>x - ip 2 ) Pi
sin (V’I - i> ) ’2

ri sin (</?! - V>x)


P2 sin (V’I - ip )'
2

r2 sin (V’I - ip ) 2

p sin (V>! >2 )’


2

and the equations of motion (10.9) take the form

rx sin (</?x - V>2) Pi n sin ( < p 1 - V>t) p 2


(T) - Fir 1 = ^ (T) (T) - + ^2 ( ? ) sin (V’I - i> ) ’ r sin
sin (V’I - i> 2 ) r 2 2 2
sin (ip 2 - ip 2 ) Pi sin ( i p 1 - y2)
F2r2 = + ^2 CO p 2 sin (V’I - V>2)'
(V’I -^2) (7.11.5)

These should, of course, be considered together with the constraint equations (1).
The quantities on the right hand sides of these equations are the generalised
constraint forces Q* and Q* . Also, £^ 1 (T ) and £^ 2 (T ) are the generalised constraint
forces and Q^ 2 -
Equations (5) can be obtained easily by means of some very simple reasoning.
Removing mentally the joint C we replace the action of the rod system A 2B 2C on
A\B\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 2B 2C.
The elementary work of the forces applied to the system A X B X C is
6 f W = F x r x 6(p x + X6 (7*1 cos tp x + p x cosi/q) + Y6 (7*1 sin ip x + p x sini/q)
= r x ( F i — X sin ( p x + Y cos t p x ) 6( p x + p x ( — X sini/q + Y cosi/q)
/
6 ipx

and the equations of motion become

£<px{T) = 7*1 (Fi - Xsirupx + Y cos(px),


£jp 1 (T) = p x (-Xsin'ip1 + Y cosi/q).

By analogy we obtain for the system A 2B 2C

£<P2 {T) = 7*2 (F2 + Xsliup2-Y cos ip2),


£^ 2 (T) = p 2 (X sin xp2 - Y cos x p 2 ) .
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 problem 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 O and A, see Fig. 7.6. A flexible
inextensible closed string passes through these rings as well
as a fixed ring 0\. 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\0A and the body moves parallel to the
vertical plane. Let the total length of the string be L, the
distance OA between the rings Z, the body weight Q, its
moment of inertia about the axis perpendicular to the plane
of the motion and passing through point O be ©o-
The system has two degrees of freedom as the body
position can be determined by the position of the pole O of
the axes Oxy fixed in the body. To this end, it is sufficient to
know the length s = 00\ and the angle <p shown in Fig. 7.6.
Indeed, the lengths of the triangle sides OA — l and 0\A = L — l —
s are known and we can construct this triangle. This
construction defines the body position and thus the angle 'ip.
Noticing that the angle of vertex O is equal to TT/2 — (<p + ip) we
or
F (s, ip, if) = 2s [L - l — l sin (cp + ^)] - (L - 21) L = 0.

(7.11.6)
This is the constraint equation relating the quantities 5, ip, (p to
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 = 1 [mvl + 0Ou>l + 2nuz (v0yxc - v0xyc)\ , (7.11.7)

where uoz = ijj 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 O are

£o = scos(y9, rj0 = — s sin ip,


we find
£0 = cos (p — sip sin ip, f/0 = — s sin (p — sip cos ip
and then
VQX = £0 cos ^ + r)0 sin ^ — s cos (ip + tp) — sps'm (ip + %p),
voy = —£>0sm'ip + 7]0cos'ip = — s s m ( ( p + 'ip) —
s(pcos((p-\-!p).
Expression for the kinetic energy takes the form T ~ \ j777^2 +
2
ms2 ip + 0O^ — 2mip [sa (a) + spa' (cr)]| ,

(7.11.8)

where, for the sake of brevity


a (a) = xc sin a + yc cos cr, a' (a) = xc cos a — yc sin a, a = ip + 'ip

(7.11.9)

and a" (cr) = —a ( a ) . The potential energy is equal to


The calculation yields
*• *2
£s (T) = m s — a (a) ip — a' (cr) 'ip —
sip2
£v (T) = m £ 4, " 2 •• •2
s ip — sa' (a) ip + 2sip + stp a
(a)
(T) = m
~ sa (a) — a' (a) sip — 2sipa' (a) + sa (a) ip2
(7.11.10)
348 7. Lagrange’s differential equations

and furthermore
Qs = mg cos p, Qp = -rag sin <^, = —mg (xc sin ip + yc cos xp) .
(7.11.1
1)

Differentiating the constraint equation (6) with respect to time

yields s [L — l — l sin (p + -0)] — s/ ^ ^ cos (p + ip) =

0.

Using eq. (6) the above equation becomes

2g2 ^L I ^ L - (<P + V>) cos (y> + = 0-

5 s~\ (L- 21) L [ip + xp'j cos (p + ip) + (p(7.11.12)


+ ^ sin (<^ + -0) = 0.
2s2 s l (7.11.13)

From eq. (12) it follows that


d ip (L - 21) 9^
L -1 . (7.11.14)
2
ds 2s l cos (p + ip)' dp
By virtue of eq. (10.9), considering ip as a redundant variable,
we obtain
£s (T) — Qs — [£*P (T) - Qtp \
(7.11.15)
,

2szl cos (p + ip) Sip (T)


(L =- 21)
Qp +L
S^p (T) — Q^p.
These equations of motion should be considered together with
eqs. (6),
(12) and (13). We notice that

Q?p - Qip = ~m9 (xc sin ip + yc cos ip - s sin p) = -mgr\c = ra^1,


(7.11.1
that is, this difference is equal to the moment of the weight
about axis 0\Q.
The latter equation in (15) thus expresses the angular
momentum theorem
Si,{T)-ev{T) = k°\ (7.11.7)

where K®1 stands for the projection of the angular momentum


about point (7.11.18)
O on axis OiC- Due to (4.8.10) it is equal to
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
(7.11.19)
where ret and rcq denote the projections of vector rc = 0(5 on
axes Oi^rj
rCt = xnow
Applying c cos the
^ - 2/c sin formulae for
above sin?/? + yc cosofi/?.
projections the pole
velocity and using notation (11) we obtain

Having calculated the time-derivative we arrive at relationship


(17) in which £^j and £^ 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 integral of energy
T+n=ft,

expressions for T and n 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 <7i, ...,gn- 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 coordinates 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 redundant coordinates qn+I , •• • , qn+m, the
constraint equations (1.4.8) having the maximum simple form
Frt+1 — qn+1 — 0, . . . , Fn+m — (JVi+ra — 0. (7.12.1)
This approach was applied in Sec. 6.5 while considering
static problems. Using eq. (10.4) we have
350 7. Lagrange’s differential equations

and by virtue of eq. (10.12) the generalised constraint forces


are as follows
Q: = ^^)-^n+r&n+r,s — 0 (^ — 1? • • • 5
5 (7.12.3)
Qn+ r=~
1
TTl)
l
-^n+Z ( / = ! , . . . , .

The equations of motion (10.1) take the form


1,... ,m). /
£ s (T ) = Qs (s =
^n+r ( - ^ ) Qn+r “ 1“ ^n+r “
= (7.12.4)

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? Qn+r — Qn-\-r — O 5 (j* — ! ) • • • ) ^ ) • (7.12.5)
In the case of stationary constraints the equations of motion
in the explicit form have the form of eq. (4.6), i.e.
n+ra n+m n+ra

^2 ^2 [CT’ F = Qs, (s = 1, • • • , n ) , (7.12.6)


<7 — 1 <7=1 p= 1

n+ra n+m n+m

^ ^ ^n+r,<7<?<7 T ^ ^ ^ ^ P\ Tl T T\ QaQp =
Qn+r T An_|_r, (v — 1, . . . , TTl) .
<7=1 <7=1 p= 1
(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 nn
^2 + ^2Y1t*7’ F, s]° QaQp = Qs (s = l , . . . , n ) , (7.12.8)
<7=1 <7=1 P—1

y An+r,<rti<r
<7=1
+ EE [a, p-n +
<7=1 p= 1
r}° qaqp = Q^+r + Xn+r, ( r = 1,... ,m).

(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
^ n+m n-\-m ^ nn
T
- SE AgaQsQa 2 2EE AgcrQsQcr
(7.12.10)
——
S= 1 <7 = 1
III IV ^ 11 vmm
11 v

EE
+ r=l <7=1 An-\-r,crQn-\-rQ<7 “I” ^ EE
r= 1 t=l

where coefficients are functions of all variables qi, ...,gn+m. The


generalised 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 8qi,...,
8qn+rn
6'W = J2Qst<ls + Y,13n+r6q„+r. (7.12.11)
s=l r= 1

The kinetic energy T° of the original system is obtained from


expression (10) at qn+r = 0 and qn+r = 0 (r = 1,..., m)

nn
(7.12.12)
T
° = oEEt».
S — l <7=1

Assuming gn+r = 0 in the expressions for the generalised


forces we find their values in eqs. (9) and (8). <3i,...,<3 n 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
n nn /
dqP dA% 9 A I P
+ q<r<ip = Qt
<7=1 <7=1 p— 1 \
d q a dqs
(7.12.13)

The square brackets in eq. (8) are given by

1 /dAS(J dAsp _ dAap\


[a, p; 0
s] 2 V dqp dqa dqs ) ’

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 differentiation. 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).
( d A f j ^ n -| 9Aap \
-1 _r
2' V 9qp dq<r 9qn+r )
'9K,n+r , dA%n+r { 9Aap
1
2 dqP dq„ \ &q[n+r
A°n ^12 * ■
A- ^22 * ■ • ^2n
21

Ah ^n2 *' ' • nn

(7.12.14)

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 g n+r. The values of An+r^
should be known only at qn+r = 0, that is only A^+r a are needed
and the coefficients An+r^n+t (^ 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

«. = - E £ { "A}0<MA+£ «)° <71 2 1 5 >


p= 1A=1 ^ ) p= 1

where the braces are calculated for the matrix of the


coefficients of the quadratic form (12) and ( A ~ p ) denotes the
elements of the inverse matrix. Substitution into eq. (9) for the
generalised reaction forces of the removed constraints yields

^n+r
~ Tl A
n+r,a T! T! { \ } 9pQ\ + Tl A
n+r,a (+p) Q° P
+
<7=1 0=1 A=1 ^ ' <7=1 0=1

^2 ^2 P
> n
+ r]° - (7.12.16)
Qn+r■
<7=1 p=l

We introduce the notation


r,AA^p)0 = M
n+r (r = 1,... ,m;p = 1,... ,n).

(7.12.17)
These values are calculated as follows: in the determinant of
the quadratic form

(7.12.18)

the p — th row is replaced by the following row


o
A n+r, 4°± . ,A 0 (7.12.19)
15
^ n-\- n+r,n*
7.13 Geometrical interpretation of the generalised constraint forces 353

Dividing the result by |A°| yields M^+r.


Recalling values (4.4) of the braces we have, due to eq.
(17),
n rN0nn
^n+r,cr 4° (4r)° [P, A; r]° = J2 Mn+r [p> ^ r]° .
<7=1
U -SE
n+r,cr r=l
(7.12.20)
Substituting these into expressions (16) we arrive, after some
changes of indices, to the equalities

An+r — EE <lAp \ W, P;n + r]1 A;r]° +


<7=1 p= 1 t r= 1 J

n+r (r = 1,... , m ) . (7.12.21)


r=l
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 A n+i,..., An+m
separately by introducing the redundant coordinates one by

7.13 Geometrical interpretation of the generalised


constraint forces
An example of the fruitful application of the Riemannian
geometry to dynamics 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 R n with the quadratic form T° ( d t )2 corresponds to the
system of material points under consideration. The covariant
components of the metric tensor on this manifold are denoted
as aa/3. The manifold R n is a part of manifold Rn+m with the
quadratic form T ( d t ) 2 . This means that the points of Rn+m with
the following prescribed values of the generalised coordinates
9n+1 =<$+\... ,qn+m = q+m
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
354 7. Lagrange’s differential equations

kinetic energy of the system R +m n

T {dtf = i (g 0dq dqP + 2g ,


a
a
a n+sdqadqn+s + gn+s,n+kdqn+°dqn+k)

(7.13.
2)
(summation over Greek indices from 1 to n and over Latin
indices from 1 to m is assumed) we arrive at T° which is the
kinetic energy of the system. Therefore
( g a g ) o = aa0- (7.13.3)
The force corresponding to the representative point of the
manifold R + is given by the vector Q+A with the covariant
n m

components Qa and Qn+S + An+S. The elementary work of the


given forces in the virtual displacement of the particles of the
system Rn released from the constraints is
QaSqa + Qn+sSqn+s, (7.13.4)

whereas
A n+sSqn+s (7.13.5)
implies the elementary work of reaction forces of the released
constraints. Newton’s second law, for i?n+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 R +m is given by n

V = r aqa + rn+s<j"+s.
Its time-derivative is
V = r aqa + r a0<f<f + 2r a,n+sqaqn+s + rn+sqn+s + rn+sqn+s.
However at points of Rn+m belonging to Rn we have, due
to eq. (1),
qn+s = 0 qn+s = 0.

(7.13.7)
Thus,
(v)o =
Pa<ja + (r<*0)qaq13,
where pa = (ra ) 0 denotes the basis vectors of Rn. Using the
relationship
(B. 13.21), we replace ( r a p )0 by its value and arrive (7.13.8)
at the
equality
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

(w)a = P7 (V + qaq^ = p,w\

(7.13.9)
The second group determines the vector
Since
rvn+t _ n
F7 — U> 0
vector wo is orthogonal to the space Rn. Vector WQ in eq. (8)
means acceleration at points Rn+m belonging to Rn. It is equal to
the geometric sum of the acceleration vector in Rn and vector
WQ 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/3 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
Q + A = (Q°)0 pa + [(Qn+t)o + (An+t)o - MZ+t (Q*)a\ *o+t ■ (7-13.11)
The first component is calculated in metric Rn and is equal to
aa(3 ( Q ) p 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

P7 (<F + + ([“’#« + *lo - M


n+t [«, # 7la) qaqPv 0+t
=
(Q )a Pa + [(Qn+t)o + (An+*)o —
^n+t (Qa)a] 0
r +
*

and the expressions for the covariant components of vector


of the constraint force
(An+t)0 = «“/([«»fcn
+ t\0-M2+t[a,(3-,7]a)
~ [(Qn+t)o A^n+t (Qa)a] (t = 1, • • • , m) (7.13.13)

which is orthogonal to the space Rn. Taking into account the


changed notation we arrived at expressions (12.21). This new
result is instructive as
356 7. Lagrange’s differential equations

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 specifying 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%N.

7.14 Application to planar systems of rods

7.14.1 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 O 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

T° = \®oql n° =
(7.14.1)
where ©o denotes the moment of inertia of the pendulum about the rotation 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 q% as well as the angle q\ + #4, see Fig. 7.7, as the parameters
describing the position of the system. In the original configuration q^ = 0, #3 = 0, q± =
0 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\ and C2 of
bodies 1 and 2 lie on a straight line OC passing through point O, 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

+ @c2 (qi + Q4)2 + m2v&2 ,


(7.14.2)

where OQ denotes the moment of inertia of the pendulum 1 about the rotation axis
O, 0c2 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 wc2 denotes the
velocity of point C2.
The coordinates of the latter point are

%c2 a cos qi - q2 sin q\ + q% cos q\ + c2 cos (qi + q4),


Vc2 a sin qi + q2 cos qi + q3 sin qi + c2 sin (qi + q4).

Thus,

v
c2 Qi +
*02 + 2/02= (a + c2)2 + 2 (a+ 02)93 2C2 (a + c2) 5154 + 2
(a + c2) q\q2 + ..

where dots imply terms which are out of interest as they contain the terms that are
of order higher than the second. We obtain

T — — {[@0 + 2m2 (a + c2) #3] q\ + 2A^2QIQ2 + 2Aj4gigr4 + . . . } .


(7.14.3)

Here

0o = © o + +m2(a + c2) 2

denotes the moment of inertia of the pendulum about the rotations axis O and

4?2=m2(a + c2), +?3 = °> +14 = ©C2 +m2c2(a+c2).


The potential energy of the system is given by

n = -G\xc1 G2XC2 = —G\c\ cosgi


G2 [(a + q3) cos qi - q2 sin qx + c2 cos {qx + q4)\
= —Gc cos qi — G2g3 COS qi + G2 (g2 + c2q4) sin </i+,... ,

where

Gc — G\C\ + Gr2 (c2 + u).


358 7. Lagrange’s differential equations

Thus

Qi = —Gc sin c/i, Q% = — G2 sin qi, By <2° = G2 COS qi, Q° = -G2 C2 sin qi.
formulae (12.17) we have

= Mj=o, @c2 + ™c2 (a + c2)

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

G2 m2(a + c2)c2
A2 G sin qi
J3 BO
- [ra2 (a + c2) + G2 COS <71] ,
A3 G2 m2c2 (a + c2) + ©c2
c2—— c------------------ ----------------
A4 n
Gf

In this example A2 and A3 are the transverse and axial forces, respectively, and A4
the bending moment in the cross-section SKS. They are the generalised forces
corresponding to the generalised coordinates < 72, #3,#4- 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+\ 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 and Vector is directed from Ai to Ai+1
and has angle cpi with the fixed axis A\x. The position of the centre of
gravity Ci in this coordinate system is described by the position vector

= s\e\ + 4e*2.
The velocity vectors of the joints are related to each other by means of the
following recurrent equalities

Vi = Vi_i +
Vi = 0 (i = 2 , . . . , n ) (7.14.4)
7.14 Application to planar systems of rods 359

FIGURE 7.8.

where Z*_i denotes the length of The kinetic energy of the i — th


member is calculated by eq. (4.7.7) as

2 Ti = rriiVi + 2mjVj • (k Cx r^) +


= mwl+ 2mi(pi(s\vi-e\-+<di<pl, (7.14.5)

where ra* and 0* denote the member mass and its moment of inertia about the axis
of the joint respectively. The unit vector perpendicular to the plane of the chain is
denoted by k, so that kCpi is the angular velocity vector of the i — th member.
Thus, we have

Tl = !©!¥>?,

T2 = \Q2& + 1m2ll<Pi + m2h<p2(p1 (42) COS /?21 - s{2] sin/?21) ,

X3 — 2®3<^3 + 2m3 ^2^2 + 2/I/2^I^2 COS @21) + m3(P3 x

h<Pi (s[3) cos/331 - 43) sin/331) + l2<p2 ^3) cos/?32 - s^3) sin/?32)

etc., where the difference of angles is introduced as follows

ft sk s ^Pk' (7.14.6)
Fk(Vi,---,<Pn) =0 (k = 1 , . . • ,n - 1 ) (7.14.12)

can be assumed to be resolved in <p2, • • • , (fn

l
Ps = fs(iPi) (s = 2,...,n). (7.14.13)
Then

<Ps = f's (^l) <Pl (7.14.14)

and the expression for the kinetic energy (7) can be written in the form

10 = \Q*(<Pi)<P2v (7-14-15)

Here

©* (<Pi) = 71°, + 2 £ < / ' + £ £ A"sfkfs (7.14.16)


s=2 k=2 s=2
7.14 Application to planar systems of rods 361

is referred to the moment of inertia of the mechanism with respect to axis Ai of the
driving member. Zeros indicate that the quantities are calculated for the values of
(p2, • • • > defined by the constraint equations.
The elementary work of the active forces (11) becomes

S'w =(QI + J2 6Vi = (7.14.17)

mz being the torque for the driving member. The equation of motion for the
mechanism can now be written down as follows

£i (T°) = 0* (^i) 0! + = m z.
(7.14.18)
Let us mentally remove the constraints, then instead of relationships (13) and (14) we
can take

<ps = fs (‘pi) + Qs, <i>s = fs +Qs, (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 <p3 and (ps by their expression from eq. (19). The
result is

A A°ls+E
dAIs fs<Pi +
ls<Pl<Ps qr
=2 V P r
dl

MstPk
(k,s = 2, . . . , n) ,
qr
fkfsV 1 + Aks (fkQs + fs4k) <Pl,
Vs
(7.14.20)

where, in accordance with eq. (8)


o dAks\° ( ^ (dAis\° (dAis\
Qs — Qk) 5 2^1
W. sk
( d<pr ) Qr = ( d(3sl, )
q8- r=2
(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 +

n nn
2<^i ^2 A 1 S Q S + ^5Z-4fcs/fc<?s ■ (7.14.22)
362 7. Lagrange’s differential equations

While deriving this equation we noted that

2
^2Y1A0^ + /«&) = 53
A
°kSfkQs
s=2 k=2 s=2 fc=2

and
Expression (11) for the elementary work should be taken as
n n
8'W = Q\8Vl + Wvi + Q') = S
(7-14-23)

s=2 s=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 = Cp\ ([1,1; r]° - Ml [1,1; 1]°) + M'rmz - Q°r (r = 2, . . . , n).


(7.14.24)
The coefficients of the kinetic energy are reduced to the form

n
M. == A°ls + J2dAksfL-
in

k=2

Therefore,

n
0
fftAu \°
/ kr \ pf pf

(dAlry ^
WJ u £ f'rfk-
V dPrk J
Taking into account that

dA°lr _ fdAlr\° dA°kr _ (dAkr\°


( / ; - 1) , tv, v»w </; ft>
d<Pl v dpr 1

we have
0
_ \(dA°kr \ f,2 _ A0 ft!
r-i i i0 f dAir
lM;rl
=-{W ,7 ^2[UPrk) f k kJk
\
7.14 Application to planar systems of rods 363

n ! no
I1’1;1] =77- - M
r
A
lr + X]
2 dip1 e* (Vl) k=2
The generalised force of the r — th constraint is equal to
o
\ *2 dA 1 r t\^)r:~AW +
\r — d/3ri +£■

Id© 7 A
1 r + £ Akrf'k +
mz
^lr + £ ^krfk^j ~Qr-
ir
2d
vi r 1 0
*(^)
Q* (Vl)
k—2
(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
h'

siiup2 — /isin( ^ l 5 ( ^ 3 = 0 » = T2 (7.14.26)

It follows from these equations that

/2 cos ip]_ = fs = /3=0. (7.14.27)


COS p2 5 COS3

The above expressions for TI,T2,T3 yield the following coefficients

An = @i + (m2 4- ms) li, A22 = ©2 H- m^l2^ ^33 — ©35

A12 = m2h ^s^cos/?2i - s^ sin(32I^ + ra3/i/2 cos/?; 211

*13 m3h (4^ cos /?31 - sin /?3i^ ,

^23 = m3l2 COS/?32 - sin/?32^ •


364 7. Lagrange’s differential equations

The reduced moment of inertia and its derivative with respect to are

^ii + 2^2/2 + ^22/2 5


©* (*>i)

^0* (Pi) dA\2


2 fk if2 ~ 1) + ^12/2 + ^22/2/2
(ty I df3 2 1

and the generalised constraint forces due to (25) take the form
0
A2 = -Cpx
dA 12 - A22f!f + ■
< 20* (A12 + ^22/2)
d(32iJ ' 20* (cp-f) d^
mz
+ {A\2 + ^22/2) ~ Q21
0* (Pi)
f dA\r$ \ 0 dA
A3 = -<Pi +
23 f2-A°2 f,3

\ d^ 31 J d(3,32
1
" (^13 + ^23/2) (A% + A23ti)-Ql
+ +
20* {Pi) dip1 0* (Pi)
(7.14.28)

The generalised constraint force A 2 represents the moment of the reaction force
applied to the connecting rod under the condition = 0. 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 A 2 by using the motion of
the two-member chain (9.32). The second of these equalities should be written in
the form

dA\2 .2
A12&1 + A22P2 ~ — Q 2 + A2 ,
d/32i ^

since the partial derivative with respect to can be replaced by the derivative with
respect to —/?2i- Substituting

V?2 = f2<P 1, ¥>2 = /2V1 + f&l


and replacing lp1 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 A3.
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 <74, respectively. The velocity of
this point is then

v2 = hqie\ + q3 ii + q4 i2,
7.14 Application to planar systems of rods 365

FIGURE 7.10.
where ii and 12 are the unit vectors of axes x and y, respectively. The kinetic energy
of the first rod is

Ti = \@iql

For calculation of the kinetic energy of the second rod it is sufficient to retain
terms of order not higher than the second in <73 and q4 as the higher order terms are
of no concern for calculating the generalised constraint forces. Then we have

T = — {©2^2 + m2 (liQi + 2Z1<71<74COS<71 — 2Zi<j,i<73sin<j'i) + 2m2sq2 [hqi cos (q2 -

<?i) - q3 sin q2 + q4 cos q2]} ,

where s denotes the distance between the centre of gravity and the joint A2. Now we
have

T = T° + 2Ai3qiqs + 2A.54<7I<74 + ZA^s&qs + ZA^&qA, (7.14.29) where T° denotes

the kinetic energy of the original two-rod system

T° = 1 (Anq\ + 2A + A22ql), (7.14.30)

where the coefficients do not differ from those at <73 = <74 = 0


An = ©1 + ra2/i, A22 = ©2, A12 = rri2sli cos (q2 - qi.)

The other coefficients are equal to

V453 = —7722/1 sin <71, A23 = —77225 sin q2,


A®4 = 7722/l COS^i, A24 = 77225 COS (72•
7772 li sin qi s sin q2
\A° A12 A22

m2 li cos qi s cos <72


|A°| A\2 A22
7772 An A\2
Mi =
O | h sin 51 s sin <72
A" 2
777 An A12
M2 =
W\ li cos qi s cos q2

What remains is to substitute the formulae obtained into eq. (31). We restrict our
consideration to the value of A 3 under the assumption that force F is absent. We
obtain

7712/1
A3 'W\ {An [A22 cos 51 - m2s2 sin q2 sin (q2 - 51)]

l\s2 cos (q2 - qi) cos q2) q\ - m2s


ml W\
{A22 [An cos 52 + rn2ll sin 51 sin (q2 - 51)] - m%l\s2 cos (q2 - qi) cos 51} %.
7.15 Cyclic coordinates 367

The same expression for A3 could be obtained by constructing equations of motion


for the centre of inertia of the rod A2A3

m2x% = A3, XQC = -h (qi sin qi + q\ cos q\) - s {q2 sin q2 + q% cos q2) ,

where qi and q2 should be replaced by their values due to the equations of motion of
the two-rod system, i.e. by eq. (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 generalised force Qs 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 explicitly 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 described by
the generalised coordinates qi, . . . , qn , among them there are m positional
coordinates qi, . . . , and n—m cyclic coordinates qm+i, • • • , qn- Then,
=
Qm+1 0, . . . , Qn ~ 0; Qs — Qs (Ql • • • 1 1 Qm) (^ — 1, • • • 5 777-)

(7.15.1)

and

T = T(q1,... ,<?m;,qn;t), (7.15.2)

and in the case of potential forces

L = L(qi,... ,qm\qi,... ,qn;t). (7.15.3)

The expression for the kinetic energy can be split into the components
T = r2 + Tj + To = T2* + u + r2* * + r; + ry + r0 ( 9 l , . . . , qm; t) .
(7.15.4)
A\\ • ■ • • Aim Am-\-l,m-\-l • • ■ • Am-\-l,n
A* = , A** =
Am l . . •• Amm An^m-\-1 • • A

(7.15.7)

These matrices are non-degenerate as their determinants \A*\ and \A**\ are the
diagonal minor determinants of \A\ which are positive by virtue of the Sylvester
inequalities (4.1.10).
We turn now to the differential equations of motion. Noticing that
dT
= 0 (s = 1, . . . , n — m), (7.15.8)
&Qm+s

and taking into account eq. ( 1), we split them into two groups: equations of motion
for the positional coordinates

d dT dt dT
= Qs (s = 1, . . . , r a ) (7.15.9)
dqs dqs

and those for the cyclic coordinates

d dT
= 0 (s = 1,.. , n — m). (7.15.10)
dt <9(7m_|_s

The latter equations can be integrated immediately to obtain the n — m first integrals
of the equations of motion
dT
dqim+s
= Pm+s = /?m+s (s = 1 , . . . , n - m) (7.15.11)
^ra+l,ra+l ^m+l,n

^n,ra+1 J^nn

(7.15.15)

1
^ra+5,ra+fc ___ (7.15.16)
|A*

Z\m+fe,m+.s being the algebraic adjunct of the element Am+ktm+s of the determinant |
v4**|, then
n—m m n—m
qm+r = Y (Pm+s - Bm+S) Am+r’m+s - £ Am+s,kAm+r'm+s■
6 =1 k=1 5=1
(7.15.17)
The quantities

Nm+r = Y (&»+» ~ Bm+s) Am+r’m+s,


6=1
n—m
m+r m+S (7.15.18)
MYr = Y Am+s,kA ’
6=1

are quotients with \A**\ in the denominator. The numerators are obtained from the
determinant |^4**| in which the elements of the r — th row

A m+r,m+1 •> • • . ,A m+r,n


370 7. Lagrange’s differential equations

are replaced by

A ra+1 B, ra+1 ,(3n-Bn

and respectively by

lra+l,fc? ? Ank'

We then obtain
m
Qm+r = -^ra+r ^ ^ Mm_^_rqk (v \, ... ,71 TYl) . (7.15.19)
k=1

These are the expanded expressions for functions (14). When we substitute them into
the equations of motion (9)

(d dT dqs dT\
= Q. (s = 1, . . . , r a ) , (7.15.20)
/ qm + T = fm +

the positional generalised velocities and coordinates, time t and the constant 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

iQrmQli''' •> Qmm ftm+1 ’ ' • *


" n—m

^ ^ /^ra+s^m+s (7.16.1)
s=1 Qrn-\-i---frn + r

Let us construct the expressions for the derivatives

dR dR dR
(k = 1, . . . ,n — m;r = 1,... ,n — m),
dqk ’ dqk' dfi. ra+r
taking into account that the arguments qkAk,t3m+r appear on the right hand side of
definition (1) both explicitly and in terms of functions /m+r. Then we have

m+r
dR_dT V? dT dfm+r y? £>/m+r _ dT dqk dqk ^ dfm+r dqk dqk dqk
7.16 The Routhian function 371
since, by eqs. (15.11) and (15.14),
dT
P
ra+r
d fm+r
and the sums in eq. (2) cancel out. This calculation is also valid for the derivatives
with respect to qk
dR dT
(V i ^
oqk oqk (7.16.3)

By analogy we find

dR ' ar a/„+. dfm +r


= E Qm-\-s E«
dfi ra+s dfm+rd(3,ra+s r= 1

and thus
dR
Qm+s — (s = 1, . . . ,n — m). (7.16.4)
d0 ra+s

Inserting expressions for the derivatives ( 2) and (3) into the differential equations of
motion for the positional coordinates (15.9) we obtain

£,(T) = £,{R) = - - - - = Q , ( s = 1, . . . , » • ). (7.16.5)

These equations suggested by Routh have the structure of Lagrange’s equations, 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 (L 1
15 • • • 5 fin i ^1 ‘ ’ * i ^2m) i
(7.16.6)
qs = qs (t; (3m+1, ... , C2m) (s = 1, . . . , m).

These expressions should be substituted into equalities (4). Then determination of the
cyclic coordinates reduces to the quadratures

z
-/ dR ■dt + q^+s ( s = 1 , . . . , n - m ) . (7.16.7)
Qm+s — to 00
m-\-s
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 constants. 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 reduced system (5) together with n - m
quadratures (7). It it necessary to add that the number of cyclic coordinates depends
upon the choice of generalised coordinates. For example, there are no cyclic
coordinates when the position of a particle in the central field is described by the
Cartesian coordinates 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 — E ^m+sQk I dt + q^+s- (7.16.8)


k=1 /
In the case of the potential forces, subtracting the expression for the potential
energy U(qi, . . . , qm) from both sides of eq. (1) and noticing that
dL
fim+s
J
~ Pm+s (7.16.9)
m+s —*

we introduce the Routhian kinetic potential LR as follows

LR = R - n (7.16.10)
Qm + 1- fm + r

The differential equations of motion (5) take the form


d dLR 0LR
£s (LR) = 0 (s = 1, . . . , m). (7.16.11)
dt dqs dqs
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)
9LR .
—qs ~ LR = h.
E dqs
(7.16.12)

It follows from eqs. (1) and (15.19) that function R can be split into three terms

R = R2 + R1+R0, (7.16.13)
where R2 and R\ are homogeneous quadratic and linear forms in the generalised
velocities whilst RQ does not depend on them. Thus,
similar to (2.6) the expression for the energy integral is written in the form

R2 + n - R 0 = 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
Pm+sQm+s
( <9T2** au or?* \ .
+
s=1 " dq-m'+s + dq m +s) ^ m+s
2 T£* + U + Tf*. (7.17.1)
Recalling definition (16.1) of the Routhian function and making use of eq.
(15.4) we obtain

R = T; + u + r2** + + rr + r0 2T2 * u r**


- - -

or

R = T£ - r2** + 7? + To, (7.17.2)


where the cyclic generalised velocities should be replaced by their expressions
(15.17).
In the case of stationary constraints the latter expression takes the following
transparent form
R = T*-T*\ (7.17.3)
i.e. the Routhian function is the difference between the quadratic forms T 2 in the
generalised velocities corresponding to the positional and cyclic coordinates.
However when we replace the generalised cyclic velocities in T 2* by their
expressions (15.17), where = 0 for the stationary con
straints, the Routhian function in terms of the generalised velocities (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 n—m n—m
m+s,m+rQm+sQm+r

IV IIV
IIV IV
^ --
-------------
^ra+s^ra+r 2iVm_|_r
=3£I> ra+s,ra+r
s=1 r=1 k=1

££ M^+sM m+rqkqi
l
$0 + + $2, (7.17.4)
k=11=1
i.e. T2 * is the sum of three components, namely the function independent
of the generalised coordinates
^ n—m n—m

^ra+s,ra+r^ra+s^ra+r 5 (7.17.5)
s=1 r= 1
374 7. Lagrange’s differential equations

the linear form in these coordinates


m n—mn—m

E ^m+s,m+r^m+r-^m+s (7.17.6)
k=1 s=1 r= 1

and their quadratic form


1 mm n—mn—m

*» = jEE EE ^m+s,m+r^m+s^m+r- (7.17.7)


k=1 Z=1 s=l r=l
Correspondingly, expression (2) for the Routhian function takes the form

R = (T2* - $2) + (2? - $1) + (T - $0) = R2 + Ri + /Jo, (7.17.8)


0

and, in the case of stationary constraints,

R = (T2* - $2) - $1 - $0 = #2 + i*i + ifo. (7.17.9)


We replace now iVm+r in eq. (5) by their values due to relationships (15.18). The
latter yield

/?, m+s -B m+s E


r= 1
^m+s,m+r^m+r • (7.17.10)

Using eq. (5) we obtain

d$p
= P m+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 $0 of iVm+s both in an associated form and in the linear form
n—mn—m
*o = 2 ^ £ Am+a'm+r (0 m+a - Bm+S) (/?m+r - Sm+r)
S—1 r=l 1 n
—m
= 2 E (/W, - Sm+s) JVTO+<. (7.17.12)

It is worthwhile mentioning that as follows from (5) both the quadratic form $0
and T2* have the same matrix A**, thus $0 has a positive definite form.
The coefficients of the linear form $1 are transformed by eq. (10) yielding instead
of (6) the following

^1 — (@m+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 system in terms
of the generalised velocities corresponding to the positional coordinates. To this aim
we replace qm+s and R in the expression
n—m
T = R+J2(3m+sQm+s (7.17.14)
S— 1

by means of eq. (15.18) and ( 8), respectively, using eq. (13) and the bilinear form of $0
given by the second line in (12).
We obtain

m n—m
T = T2* - $2 + Ti* + £ g k Mkm+S (Pm+s - Bm+s) + To -
k=1 s=l
^ n—m n—m / m
^ 53 (Prn+s ~ Bm+s) ^rn+s + 53 An+s ( Nm+8 ~ 53 QkM'
k
m+s
s=1 s=1 k=1
(7.17.15)

Replacing T on the left hand side by the sum

T = T2 + Ti + T0, (7.17.16)

and Ti by the following expression

— 5 v BkQk “b 5 > Bm+sQm+s


k=1 s=l
n—m / m \
= Ti + 53 Bm+s Nm+a — ^3 Mm+SQk , (7.17.17)

s=1 k=1
we arrive, after rearranging the terms on the right hand side of eq. (15), at the
following relationship

T2 + T* + E +T
°
S=1 V fc = l /
m n—m
= (T2* - *2) + T; - M*+sl?m+s + To
/c=l S=1
^ n—m
+ 9 E (A™+s + Bm+s) N,
ra+s-
s=l
By virtue of eqs. (12) and (8) this can be simplified to give

T2 = R2 + 3>o* (7.17.18)
376 7. Lagrange’s differential equations

It follows from this representation that


^mm ~ n—m n—rn "

R2 = T2* - $2 = - EE qm Aki - EE ^m+s,m+r*-^m+s‘^m+r


fc=l /=1 _ s=l r—1 _
(7.17.19)
is a positive definite form of the generalised velocities qi, . . . , <7m corresponding to
the positional coordinates. One arrives at this conclusion from the following
considerations. Let q° denote the initial values of the generalised velocities qk and let
us assume that the cyclic momenta /? m+s equal the initial values of the coefficients L?
m+S, i.e.

at t = t0 Bm+S ( < ? ? , . . . ,q0m-,t0) =/3m+s (s = 1 , . . . , n - TO) .


(7.17.20)
Then, due to eqs. (15.18) and (15.17), the initial values of the cyclic generalised
coordinates are
n—m
q°m+s = - ]T Mi+S (ql t0)(7.17.21)
k=1

Under such initial conditions <E>o = 0 and R2 = T2, that is R2 is positive. Given
that the initial moment as well as the initial values for the positional coordinates
q^ , . . . , q^ and the corresponding velocities q± , . . . , q^ 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 = n - p2*]4m+.=/ra+. = i?2 + Ri + Ro, (7.17.22)


where R2 being quadratic in <71, •••, Qm is positive definite, whilst the linear part Ri
due to eqs. (9) and (13) is equal to
m n—m
Rl = E ^ E M™+sPm+s- (7.17.23)
k= 1 s—1

Here the coefficients 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

.* n—7 11 n—7 11
R« = -So = -2 £ £ Am+s-m+^m+s/3m+r (7.17.24)
s=l 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
T2* with respect to This means that the
quadratic form T2** is an associate form for —RQ
-j f t f f t f t f f t
?2 =
2 ^m+s>m+r^m+s/^m+r = ~^0' (7.17.25)
s=1 r=1

The equations of motions for the positional coordinates are given by the form

d dR2 dR2 dRp


Qk + Tfc + dqk (k = 1, . . . , m ) , (7.17.26)
dt dqk dqk

where Tk denote the gyroscopic forces (Sec. 7.3) corresponding to Ri. R2 can be treated
as the kinetic energy and —RQ as the potential energy. According to (25) the latter is equal
to the kinetic energy of the hidden motions (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 energy R2 and the corrected potential energy of the system II — RQ
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 T2), 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 R\ =0 the gyroscopic forces are absent and the system is gyroscopically uncoupled.
By virtue of eqs. (23) and (15.18) this occurs if

Am+s,k =0 (s = 1, . . . , n - m ; k = 1, . . . , m), (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

mr = - l / ( r ) ,
378 7. Lagrange’s differential equations

where r denotes the position vector of the moving point referring to the attraction
centre O, / (r) expresses the dependence of the force on the distance. Its moment
about point O is obviously equal to zero as

rx
~/(r)= °-
Therefore

r x mr = — r x mr = 0,
at

and the angular momentum of the particle about centre O is constant

K° = r x mr.

This means that the trajectory of the particle is a curve which lies in the plane
perpendicular to the vector K° and passes through the attraction centre. We take the
latter as the origin of the polar coordinate system r, ip in this plane. The expressions
for the kinetic and potential energies take the form

T = ira (r2 + r2<p2) , II = J f(r)dr. (7.18.1)

The coordinate is cyclic and thus the corresponding integral

mr2(f = fiy (7.18.2)

expresses the known law of areas implying the constant value of vector K°. The
Routhian kinetic potential is

r*
LR = 2 - T2** - n = imr 2
-^+ J f (r) dr, (7.18.3)
ro

and the differential equations of motion takes the form

Q2
Sr (Lr) = mr----------------^3 + / (r) = 0. (7.18.4)
mr6

It admits the first integral

T2* + T2** + n = |mr2 + ^ + J / W dr = h, (7.18.5)


7-0

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 O, the pole angle $ (the complementary angle
to the latitude) and the longitude A, then we can decompose the vector v into
components along the radius VR, along the tangents to the meridian v# and the parallel
v\ such that

VR = R, v# — Rd, v\ = R\sm$. The expressions

for the kinetic and potential energies are

n
T = Im [k2 + R2b2 + R2\2 sin2 d
'), U(R) =
RQ
J f(R)dR. (7.18.6)

As A is the cyclic coordinate, then the corresponding first integral is given by

BT .
(7.18.7)
2
(3X — — = mjR Asin ^, 2

and the Routhian kinetic potential is determined as


LR = \m (ii2 + R2#2)-----------------------------=- (7.18.8)
-n.
2V ) 2mR2 sin2 Pi
&
The differential equations of motion are

Pi
mR — mR& + IT' (R) = 0,
mR sin2 •& 3
(7.18.9)
m (jR2^ — mR2 sincos •&
= 0.
3
•&
They admit the first integral

^m(R + R i)2) +U(R) , o .


2 2 p\
2 V / v +■
’ 2mR2 sin2 ■&
= h. (7.18.10)

This first integral could be written directly from eq. (16.14). By means of this integral
the quantities $ and $ 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 O on axis Oz
which is the axis of symmetry of the inertia ellipsoid at point O. We introduce ”half-
moving” axes n, n',^ (see Sec. 2.10 for detail) and take the projections of u; on
these axes as quasi-velocities. Then
setting

0° = A (nn + n'n') + Ci'3ig, (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)

T = -A (LU\ + cu%) + ~Ctu\. (7.18.12)

By virtue of eq. (2.10.4) we can write the kinetic energy in terms of Euler’s angles
ip,ft,p in the form

T = 1A (ti2 + i>2 sin2 d)+^c(v + i> cos i?)2 . (7.18.13)

The potential energy is

n = Mg(c = Mgz cos#, (7.18.14)

where (c denotes the coordinate of the centre of inertia along the upward vertical 0£
and z stands for its coordinate along the axis Oz of the body symmetry. The cyclic
coordinates are ip and p and the corresponding integrals are given by

Aip sin2 ft + C (p + ip cos #^ cos ft — (3^, C (p + ip cos ft^j = (3^.


(7.18.15)

The Routhian kinetic potential is then determined by the equality

r 1
/Vo2 -/^costf)2 P%\ „ q
LR = -\A& - W v2 --77 - Mgz cos d (7.18.16)
2\ Asm ft CJ

and the integral of energy is

Ad2 + ^ ~ /jy cos^) + 2Mgz cos d = 2h. (7.18.17)


A sin2 ft
Putting cos$ = u we come to the familiar differential equation for the cosine of the
nutation angle

Aii2 = 2 (h- Mgzu) (1 - u2) - j (^ - 0vu)2 . (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
ft remains small. Euler’s angles are not appropriate generalised coordinates as angle
ip is not small and sin 2 ft 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
n
£
cos (3
V
0
c
— sin (3
n' — sin a sin (3 cos ce — sin a cos (3
>3 cos a sin /3 since cos a cos (3

(7.18.20)

The expressions for the kinetic and potential energies take the form T = - A ^d2 + (3

cos2 + “(7 (^p + (3 sin , II = Mgz cos a cos /?,


382 7. Lagrange’s differential equations

indicating that only coordinate cp is cyclic. The integral corresponding to the cyclic
coordinate as well as the Routhian kinetic potential are as follows

Ptp = C (<£ + /? sin aj ,

LR = iA (a2 + (3 cos2 o;^ + PVP sin a — — Mgz cos a cos p.


(7.18.21)

The energy integral has the form

A (a2 + /?2 cos2 c^J + 2Mgz cos a cos (3 = 2ft. (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 0( is constant. The
external forces acting on the top is the weight and the reaction force of fixed point O
and thus the moment of these forces about the fixed axis 0( is zero. When we adopt
angles a and f3 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's

Acji = - Aa, AUJ2 = Af3 cos a, CUJS = C (^p + (3 sin = (3^,


(7.18.23)

we obtain by projecting on axis 0(

Aa sin (3 — A(3 cos a sin a cos f3 + (3^ cos a cos (3 = 7, (7.18.24)

where 7 is a constant value.


The equations of motion corresponding to the Routhian kinetic potential ( 21) take
the form

A^a + (3 cos a sin — (3^(3 cos a — Mgz sin a cos (3 = 0,

A (j3 cos2 a — 2a(3 sin a cos a'j + f3^a cos a — Mgz cos a sin (3 = 0.
(7.18.25)

When a and (3 are small we arrive at the system of two differential equations

Aa - (3^(3 - Mgza = 0, Aj3 + j3^a - Mgz(3 = 0. (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 [20].
7.18 Examples 383

7.18.3 System of two heavy tops


The first heavy top having a fixed point 0\ rotates with angular velocity co. Its inertia
tensor at point 0\ is denoted as ©i, the centre of inertia C\ lies on the symmetry axis
0\Z\ with abscissa z\. Point O2 on axis 0\Z\ is the point of support for the second
heavy top rotating with angular velocity Q. Its inertia tensor at point O2 is denoted by
©2, the distance O1O2 is / and the abscissa Z2 of the centre of inertia C2 lies on the axis
of symmetry 02^2-
We use two ” half-bounded” systems of axes introduced in the previous example,
axes n, n', ig for the first heavy top and n*, n*', ig' for the second one. The system has
six degrees of freedom and the ship angles

5 5 5 Oi2 5 @2 5 V^2

are taken as the generalised coordinates.


In accordance with eq. (19) the quasi-velocities are the projections UJ\, LJ2, UJ3 of
vector co and the projections , UJQ of vector Q on axes of the
corresponding ” half-bounded” system.
The kinetic energy of the first heavy top is given by eq. (12)

T\ — -Ai +^2) + 2^lUJ3- (7.18.27)

The kinetic energy of the second heavy top is calculated by means of eq.
(4.7.7)

T2 = ^M2V2O2 + M2VO2 • (fi X i*') 22 + • ©2 • (7.18.28)

Here M2 denotes the mass, vo 2 the vector of velocity of the supporting point and ig Z2
the position vector of the centre of inertia of the second top with the origin at point
O2. Then we have

vo2 = x Zig, — (u> • ig)2 = Z2 (a;2 + a;|) (7.18.29)


u
o2 = V
and moreover
v
o2 • (^ X ig) = Z (UJ • ig) • (o X ig ^ .

Noticing that

(jj x ig = (a;in + Cc;2n' + u3ig) x i'3 = — n'ccq + ncc;2, | (7.18.30)


Q x ig = ^J4n* +
7
x = -n*^4 + n*u;5,

we find

vo2 • x ig ) = l • n* + u z u s n • n* — • n* — Ccqcc^n'
• n*
384 7. Lagrange’s differential equations

and the kinetic energy of the system takes the form

T= - \Ai (^I + ^2) + A2 (^4 4- a;2) + C\LU\ 4- Ce^g] 4- (7.18.31)

B ^i^n' • n* + u^u^n • n* — u^u^n • n* — ,


where A[ = A\ 4- M2/2, B — M2IZ2. 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—— ^d2 + /3j COS2 + A2 ^d\ + /?2 COS2 OL2^ 4-

C\ (^p1 + /?! sinai^ + C2 {(p2 + P2 sina2) I +

B OL\OL2 (COS ai COS 0L2 4- sinaq sin a2 COS /?) + /?i/?2 cos ai cos a2 cos /?4-

di/32cosa2 sinai sin/? — d2/?i cosai sina2 sin/? , (7.18.32)

where /? = /?2 — (3V The potential energy is equal to


n = QI(C!+Q2(C2
= Q\Z\ cos OL\ COS (31 + Q2 (l cos oq cos f31 + Z2 cos a2 cos /?2).

where £Ci and Cc2 are the abscissas of the centres of gravity of the tops referring to the
upward vertical Oi(. Next, Q\ and Q2 denote the weight of the first and second tops,
respectively. Now

fl = (Q\Z\ + Q2I) cosai cos/?! + Q2Z2 cosa2 cos/?2. (7.18.33)

The cyclic coordinates are ^ and (p2 and the corresponding integrals are given by

C1U3 = Cl (</?! +/31sinaii) = 7X, C2w6 = C2 (ip2 + (32sma2^ = 72.


(7.18.34)

The Routhian function, due to eq. (17.22), takes the form

R = T* - T2** = i (aj + cos2 + A2 (a2 + /?2 cos2 a2)l +

B |4d (cosot\ cosa2 + sin an sina cos/?) + j3ij32 cosax cosa cos/?+
2 2 2

di/?2 cos a2 sin oq sin /? — d2/?! cos oq sin a2 sin (3 4- i\(31 sin a\ +
72^2 sinot2 - = R2 + Ri + R0, (7.18.35)
7.18 Examples 385

and the integral of energy is as follows

fl2 + n = fe.
(7.18.36)
We can construct another integral, namely the integral of the angular momentum
about the vertical axis Oi(. Calculation of the resultant angular momentum K.Ql about
the fixed supporting point is performed by means of formula (4.8.10). For the first
heavy top

K^1 = 0i • CL? - - A\ (c^iii -K Ct?2n/) + C0J3I3. (7.18.37)

In order to calculate K^1 it is necessary to have the expression for the vector Q2 of the
resultant momentum for the second top

Q2 = M2VC2 = M2 (u> x li'3 + n x zi3 ) . (7.18.38)

K 1 _ Zi3 x Q2 T z2il x MVO2 + 02 •


2 —
= -M2Z2 (CL?II1 + U?2n0 T ^2 ^CJ4n* + UJ5II* ^ + C2M3I3 B ^3 X ^- +

W4I1* + CL?5I1*^ +ig X (— CL?in' + Ct>2n) . (7.18.39)

Now we need the projection K^1 of the vectorial sum

KGl - Kf1 + K^1

on the vertical. Denoting the unit vector of the upward vertical as k we have

k x io3,= n' sin /T — n sin a\ cos /T ]


, 1
t (7.18.40)
k x ig = n* sin (3 — n* sin a.2 cos (3 J
2 2

and then

(k x ig) • n'* = n' • n* sin (3 — n • n* sin a\ cos /3


1 1

cos a 1 cos 0L2 sin f3l + sin a\ sin ot2 sin (32

etc. We obtain then

K®1 = — {A[ (CL?I sin/31 + U2 sinai cos/?x) + A2 (CL?4 sin/?2+


CJ5 sin a.2 COS /32) + B [CL?4 (COS a\ cos <22 sin (31 + sin a\ sin 01,2 sin j32) +
CL?I (COS a\ cos a.2 sin (32 + sin a\ sin (*2 sin (31) + a; 5 sin a\ cos /32 +
0J2 sina2 cos/?!]} + CICL?3 cosai cos/31 + CW6 cos <22 cos/32 = 7,
(7.18.41)
386 7. Lagrange’s differential equations

where 7 denotes a constant value. Under our choice of generalised coordinates this
integral is not a cyclic one. Using angles $1 and $2, ^2? ^2
for describing the positions of the first and second tops, respectively, instead of eq.
(33) we would obtain

n = (Q\Z\ +Q20cos^1 +<72^2 cos


$2,

where $1 and $2 denote the angles between the tops’ axes and Oi£. The expressions for
the kinetic energy would contain angles $1 and $2 (similar to a\ and a.2) and the
difference /ip2 ~'lPi (similar to (31 and P2 in eq. (32)). Thus, if we adopt

/1/1
^2 = X+-w,

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 Oi£ 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 c^, /?i5 a^, Pi we obtain

R
2 [^1 (<*i + ^1) + ^2 ( <*2 +P2) + 2B 6ti6t2 + P1P2} +
71/31Q'I +72/^2, (7.18.42)

n — — — ( Q 1Z 1 + Q 2I) (07 + Pi) — —Q 2Z 2 ( 0% + P 2) •

(7.18.43)
The equations of motion are
ai + bibt2 ~ s\Pi — C \ OLI — 0, Pi+ biP2 + siaq - ciPx = 0,
&2 + b^Oi 1 — S202 ~ c2&2 = 0; @2 4“ ^201 + ^20:2 ~ 2@2 ~ 0, C

(7.18.44)

where

,B B 7i 72 Qi*i + Q2I Q2Z2


bi = -77, f>2 = ~r, Si = -77, S2 = -r, Cl =-------------------------------------------------------, c2 -
A2’ A2
(7.18.45)

The equations of motion can be reduced to a system of two second order


differential equations in the following unknown variables

ai + i(3l = zlt a2 + i02 = z2 (7.18.46)


7.19 Quasi-cyclic coordinates 387

which are complex numbers determining positions of the vectors if3 and i3 in the
plane parallel to Oi£r/. These equations have the form

z\ + biZ2 + is\Zi - c\Z\ =0, Z2 T b2zi + is2z2 - c2Z2 = 0. (7.18.47)

The particular solution is sought in the form

Zi = ZieiXt, z2 = Z2eiXt, (7.18.48)

where Z\,Z2 and A are constant values. Substituting (48) into (47) yields the system of
linear homogeneous equations

Z\ (A -f- Si A
2
+ Ci) -f- Z b X — 0, 1 2 1
2

(7.18.49)
^2^2A + Z2 (A2 + S2A + C2)
2

— 0, J
having a nontrivial solution for Z\ and Z2 provided that the determinant is zero. This
condition leads to the following equation fourth order in A

(A2 + si A + ci) (A2 + S2A + 02) - M2A4 = 0. (7.18.50)

The particular solution exists for any of the four roots A&, the values of Z\ and Z2
being related to each other by the condition

%=- 2 --------------= _Afc + 52Afc + C2 (7.18.51)


^2 Ak + s\Xk + ci ^A k
The general solution of the system of equations of motion (47) can be written in the
form

3 4
zi = J2 CkbiX keiX*\ z = Y.Ck (A* +
2
2
51
+ Cl) (7-18*52)
k=1 k=1

The necessary condition under which the axes of the heavy tops conserve nearly
vertical positions is that the absolute values of zi 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 A also exists. The real part
of one of roots iA, iX 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 qi, . . . , qe 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 < / i , . . . , qm be positional and qm+i,... , qn quasi-cyclic coordinates. The
differential equations of motion are then split into two groups

£s{T) =
JtWs~Ws=Qs{qi"-"qm) (s = 15• • • ’'m)’ (7 19 1}
- -

Pm+s = Qm+s (qi, ■ ■ ■ , qm) (s = 1, . . . , n - m ) , (7.19.2)

with Pm+s denoting the quasi-cyclic momenta

dT
Pm+s —dqm+s (s — 1, . . . , n — m) . (7.19.3)

They are not constant under the motion but all formal constructions of Secs. 7.15-7.17
remain valid.
In this fashion we can resolve the system of linear equations (3) for quasicyclic
generalised velocities

qm+s fm+s {qii • • • 5 <Zra3 qii • • • 3 qmiPm+1 3 • • • 3 Pn \ t) 3 (7.19.4)

construct the Routhian function

n—m
R = r - E Pm+sqm+s (7.19.5)
. s=l Qm + r = fm + r

by means of the rule (16.1), and write the differential equations of motion in the
form

d 3R
Ss
^ = dtdf ~ ~dq = ,<3W
) (s = l , . . . , m ) , (7.19.6)

OR
= -qm+s ( s = 1,... , n - TO) . (7.19.7)
dPm+s
7.19 Quasi-cyclic coordinates 389

The systems of differential equations ( 6) and (2) should be considered together.


Determining the positional generalised coordinates, the corresponding generalised
velocities and quasi-momenta, we find the quasi-cyclic coordinates 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 = r- <:= Rz + R, + R», (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 <71, . . . ,<?m. This follows from expression (17.19) of this form
and rule (15.17) of construction of M^+s. The term

Ri = EE Mm+sQkPm+s (7.19.9)
k=1 s=l

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

1 n—m n—m
fl o - j E E Am+s’m+rpm+sPm+r. (7.19.10)
s=1 r—1

Thus, taking into account eq. (2) we have

m n-m 'QMv 8M^+S


n—m
\
-£k(Ri) = E^E __fc „
V=1 s= 1 Pm+s - EMm+sQm+s
„=1 = s =1 v 9 9 ^
n—m 7=1
Tfc — 'y ' M^n+sQm+s
(7.19.11)
S= 1

where T& denotes the generalised gyroscopic force. Equation ( 6) can be written as
follows

£k {R2) = Qk + r& — ^ + dRp


(k = 1, . . . ,m).
dqk
s—1
(7.19.12)

We consider now the case of the potential generalised forces correspond-


ing to the generalised positional coordinates
0 __ an

Wk o ( k = 1,... ,m). (7.19.13)


dqk
390 7. Lagrange’s differential equations

In this case equalities (5.3.12) yield the following

dQk = dQm+s
(7.19.14)
dQm+s &Qk

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 = > ----( n—m


dRo dRo . , ^ ORQ .
dt
dqk qk . dpm+sPr

we repeat the derivation of Sec. 7.2 with respect to the differential equations ( 12) and
arrive at the equality

7 7Tb 71/ 7Tb ft fll/ ^ -j—y

(R2 + n - i?o) = - £ qk Y, Mi+sQm+s - Y. J-^Qm+s


Tn s
dt k= 1 s= 1 S=1 °P ^

d (Ri + RQ)
- - Ek~l dpm+s Qm+S' (7.19.16)

As i?2 does not depend upon the generalised momenta we have

d(R!+Ro) _ dR
Qm+si (7.19.17)
dpm+s dpm+s

and relationship (16) serves to determine the power of the non-potential forces Qm+s
in the system with quasi-cyclic coordinates

A” — ^ ^ Qm+sQm+s — ^ (-^2 II -^o) • (7.19.18)


s=1
8

Other forms of differential equations of


motion

8.1 The Euler-Lagrange differential equations


These equations differing from Lagrange’s equations by introducing quasivelocities
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 systems 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 6T on the right hand side by the expression


A. I. Lurie, Analytical Mechanics © Springer-
Verlag Berlin Heidelberg 2002
392 8. Other forms of differential equations of motion

n
{ BT BT \

Substitution into eq. (6.4.17) yields


BT
d dT v BT _

S=1 S=1 BL
U
A dT £ A a r £ ^ dT r.c . .

s=l
C/CJS
_ ^ I/
s=l
07TS
0 I/
s=l
-*■ o ' /
s=l
nnn
EEE^^.-EE^-

<« ■ »>
r=l t=l s=l r=l s=l

Cancelling out the underlined terms we arrive at the relationship


d dT .ST ^ r ST ST „
_ _ ____ n n = 0.
E^ dt Boo* + EE^ST"* + 2><75T - sr - p-
S=1 r= 1 t=l r=l (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 67TS 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 BT BT dT dT dtor dirs
dt BLUS + EE*sz" , + E £ Ps (s = l , . . . , n ) , (8.1.5)
r=l t=1 r= 1

The kinematic relationships (1.5.25)


n+1
qs = ^2bskuik (s = 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 2 n first order ordinary
differential equations in the same number of unknown variables

CJI,... ,qn. (8.1.7)

The equations simplify and take the form

d BT r BT BT ,
p
+ EE^.aS;"<-a^ = * ( 8. 1. 8)
dt 8ujs
8.1 The Euler-Lagrange differential equations 393

qs = ’Y^bskUk (s = 1 , . . . , n ) , (8.1.9)
fc=1

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 linear
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
U)s = ^askqk +as,„+i = 0 (s = l , . . . , l ) (8.1.10)
fc=1
if we use eq. (1.5.22). Then we have
n

Sirs = '^2askSqk = 0 (s = l , . . . , / ) , ( 8. 1. 11)


fc= 1

with l being the number of non-holonomic constraints. The summations over s and t
in eq. (4) should be performed from l + 1 to n

nn p.rp n dT dT
d d T dt
E 6n* +E E + dwr dirs
-Ps = 0.
s=Z-t-l duo s r=l t=l+1 r=l
(8.1.12)
As variations 6irs for s = l +1, . . . , n are independent, the expressions in the brackets
must equal zero. We obtain n — l equations of motion

d dT dt dT r dT dwr £s dirs
Ps (s = l + 1 , . . . , n ) ,
dtos duor
r—1 t=l+l =1
r
(8.1.13)

whose number coincides with the numbers of degrees of freedom. The fol lowing n
kinematic relationships
n+l
b
qs = E ^k (s = 1 , . . . ,n) (8.1.14)
k=l +1

need to be added to these equations. We have altogether 2 n — l first order equations


in the same number of the unknown variables

CJ/+I, ... ,uon;qi, . . . ,qn. (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


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

o ; i , . . . ,ui, (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 CJJ+I, ... ,un. 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 us are the generalised coordinates

us = qs (s = l , . . . , n ) , (8.1.17)

since, as pointed out in Sec. 1.9, all symbols and ers 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 7^. 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 + 1, . . . , 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 Lagrange’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 performed 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
d dT dT
dt dip dip = m3,

d dT dT
dt an" ~dd = ™N

d dT dT
= ^3,
dt d(p dtp

if the right hand sides are given by eq. (5.2.9) we would obtain cumbersome
expressions with no symmetry. By using the expression for the kinetic energy 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

@1^1 + 2
(03 — © ) ^2^3 =
1

©2^2 + (©I ~ ©3) ^3^1 = / (8.1.18)


©3^3 + (©2 ~ ©l) ^1^2 = m
35 J
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

7 o? 2 + . . . , ( 8. 2. 1)
T=-M ----- (^1 T ^2) T T 2(2CJ2^4 — 2(2CC?ICC?5
0 o
where the terms CJ\ and UJ\ 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 , . . . ,
I and in eq. (1.13) by the indices Z + l , . . . , 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

and
2 o
pi — -Ma w3, p\ = Mau2, pi = -Mauoi. (8.2.3)
o
Recalling that Cb\ , d>2, ^3 denote projections of the angular velocity vector UJ on
axes Oxyz fixed in space, we obtain expression for the elementary work of the active
forces applied to the sphere in the form

SfW = V • <$r0 + m° • 0 — ViSxo + VzSyo + rh° Sni + m°87T2 + rh°87r3.


However by eq. (2.10.19)
8x o = a8n2, 8yo = —afiifi.

The generalised forces corresponding to the quasi-coordinates are

Pi — rhi - a\2, P2 = rh° + a Vi, P3 — (8.2.4)


The first Euler-Lagrange equation is given by

I "-1
It is necessary to specify the non-zero indices with right subindex 1. They
are

731 = !- 721 = -1, 731 = -«•

We obtain
dT dT „ dT
—Ma Cb\ -7——— -7——Co2 — 777—— Pi — 0.
5 ~ A ' du2 du)■ du)A

By virtue of eqs. (2) and (3)


dT „ dT „ dT „ „ ~ /7 2 \~~
x?* ~ ^ ~ 35T"3 = M“ (5 ^ 5 “ = °'
the Euler-Lagrange equations for quasi-velocities C02 and 0)3
are derived by
analogy. Finally we obtain the following three
equations
7 • ~ >
-Ma o>i= raf — al/2, 2

5
7 • ~ (8.2.5)
-Ma2 0)2= Ta° + aVi, >
0
2
-Ma2 CJ3= 171°.
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

Ui, LU 2 5 £>3, ^ P, X 0 , 2/0-

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

T = -M (±Q + yl) + - - M a 2 (fi2 + ip2 + p2 + 2ippcosfisj .

Equations for the non-holonomic constraints are

#o — a (& sin ip — p sin $ cos ip\ = 0,


Vo + a, (it cos ip + p sin $ sin ipj = 0.

In order to avoid mistakes in numbering coefficients a^s we construct the


general equations of dynamics (6.3.4) which are as follows

M (xoSx0 + yofyo) + -Ma2 + ipp sin Sd +

-Ma2 (ip + ip cos'd — pdsind'j 8ip + j^Ma2 + ^Pcos^ ~ ipd sind^ 8p — V\8XQ — V28y0 — m^8ip — m

%8d - m°8p = 0. (8.2.7)

Equations for the non-holonomic constraints yield

8x o — a (sin ip8d — sin d cos ip Sip) = 0,


8yo + a (cos ip8d + sind cos ip8p) = 0.

Multiplying these equations by the constraint multipliers —Ai and — A 2, adding the
results to the left hand side of eq. ( 7), and equating the coefficients of variation of the
generalised coordinates we obtain five equations of dynamics

Mxo — Ai + Vf, Myo — \2 + V2,

-Ma2 ^ + ipp cos d'j = — a (Ai sin ip — \2 cos ip),

-Ma2 (ip + ip cos d — ipd sin d^j = m® + a sin d (Ai cos ip + \2 sin ip),

-Ma2 (ip + ipcosd — pi}sind^ = m®.

(8.2.8)
|Ma2 1 + ip(p cos'#] + Ma(xosin^ — fjo COS ip)
= + a (Vi sin ip — V2 cos

|MO2 1 {^p + ip cos # — iptf sin # j — Ma (XQ COS ip + yo sin ip)


= m3 — a sin # (Vi ip),
cos ip + V2 sin
^Ma2 1 (ip + Ip cos # — (p$ sin = rh®.

(8.2.9)

The two equations (6) for the non-holonomic constraints should accompany 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 coincidence 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 UJ4 and UJ
5 is given by

+
T=-M 2^ — 2aujs^4 — 2aujiu;l sin$ . (8.2.10)
2
Here, due to eq. (2.10.14)

LOI — $, L02 = ^ sin$, CJ3 — (p -\- ip cos $,

and equations for the non-holonomic constraints (2.10.26) take the form
cul = LU4 + (ILJ3 =0, CJ5 = CJ5 + auj\ siniJ — 0, (8.2.11)
where 004 and 005 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. Accounting for the weight
we obtain the following expression for the potential energy

n = Mgz = Mga sin i9. (8.2.12)


Since in this case 7Ti = $ we obtain

on
—Mga cos $, P2 = P3 = 0.
/ 8T\ 30 dT 1 w ,
71— = -Ma2ixj2 OLJ 2

IIto *^3
Pl
2

* *
Pa = 2Ma2o)3, p* = -Maw3, lh = —MQJUJ\ sini?.

The derivatives
dT dT dT
dixi ’ diX2 ’ dir3
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)

721 = -cottf, 721 = 1, 712 = cottf, 712 = (8.2.13)


4
T52
1
• 0 ’ 0^32
5 1
CL
• Q 5 T42
5
• G ’ 0^23
g CL
•Q’
sintf smv smv smv
We obtain the system of differential equations
-UJl — -Uj\ cot d + 2CJ2<^3 = — - COS I?,
Z Z CL
1 1 (8.2.14)
-cj + ^
2
cot =

2cj - W1UJ2 = 0.
3 ,

If instead of t we take 1? as the independent variables, then Vs = v'su

1 (s = 1 , 2 , 3 ) ,
as $ = CJI and the prime denotes a derivative with respect to $. The second and third
equations (14) take the form
LU'2 + L)2 cot {} — 2cj3 = 0, 26^3=6^2* (8.2.15)

Removing (J2 we arrive at the equation

u'l + 6J3 cot — CJ3 = 0. (8.2.16)

Under the initial conditions

i = 0, # = tf0, W i = a ; ? , W2=^ = 2^°, (8.2.17)


the general solution of the differential equation (16) integrated by means of the
hypergeometric series has the form
^3 = F {&,$(), , (8.2.18)
from which a;2 is obtained by differentiation. Now the first equation (14)
reduces to the form
3d 2 Q 19n
^1 = — COS 17 + -UJ2 cot v — 2CJ2<^3,
4 dr/ CL 2
yielding CJI (1?) as a quadrature. A further integration determines $ (£), [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 $ and (ps are given
by eqs. (4.13.13)-(4.13.17). Due to (1.13) the equations of motion are

d dT r dT
dT U)s Pi,
dirr
dt duj ^'s7 dur

d dT A A r dT_ dT
+ Us dir8 Ps-
dt du8 “ “ 7s8
dur

Recalling equalities (1.10.15) we obtain

d dT dT dT
dtdur dui^ dnr 8

7
(8.2.19)
d dT dT dT _
dtdu dw\Wl
8 dir8 8

Now we need momenta p\,Pj,p% and then to take into account the equations for the
non-holonomic constraints

U\ = U>2 = U3 = OJ4 = UI5 = U>Q = 0. (8.2.20)

We obtain
dT \ .
-x— = P1U1 sin x cos x + u8v cos x,
v^l/0
' ffjp \
-x—) = (/ ■ * +Mi sin 2 xA+uisi'sin x, (8.2.21)
<dur J0
f dT \ , .
a— = vtus + vu sinx 7
\ du>8 ) 0

implying that the quasi-velocities o; 2, • • •, CJ6 can be taken to be equal to zero. With
the help of eqs. (1.5.17) and (1.10.14) we find

dT A dT BT n
=
A--- — / v 047TT" — 0,
d(
0*7 “ ir OX
> (8.2.22)
dT 8 A QT QT
=y^5r8__ = __ = /ijai sin x cos x "I IAJJJU8 COS 2 -

dqr d\
r= 1

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

(/x + p1 sin2 x) sin x cos x + vu8 sin x = Pi, 1


vlu8 + t/ui sin x + VU7U8 cos x = Ps- \
(8.2.23)
8.2 Examples 401

Denoting ujg = X,LJ7 = V, where V is the absolute value of velocity of the joint B
linking the axles we rewrite these equations in the form

(M + MI sin2 x) V + Mi^Xsinxcosx + ^ X s i n x = P 7,
(8.2.24)
X + jV sinx
vl

As follows from eq. (4.13.20), the value vl presents the moment of inertia of the
front axle reduced to the axis of joint B, Pg6x denotes the elementary work of the active
forces in the virtual displacement, determined by variation of angle y, and the
generalised force Pg 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, P 7 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 = Pg = 0 and the initial conditions are
given by

t = 0, E = E0, X = Xo> X = Xo- (8-2.25)

The differential equations of motion admit two first integrals. One of them

X+yVsmx = Xo + 7^ o s i n x 0 (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

(ji + fix sin \) V + 2 vj(V sin x + l^X2


2 2

= (fi + fii sin2 Xo) Vo + 2^X0sin x0 + ^Xo-

This equality can be recast in the form

M + (/^i - 7) sisin 2
%
V'z + 1'l[x + y sinx

^o2 + vl[ Xo + y sin Xo ) • (8.2.27)


sm2 Xo

Equations (26) and (27) yield the formula relating velocity V with angle \

V_ H+ 0»-i) sm Xo
Vo 2
(8.2.28)
\ M + (MI - y) sin X
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
O 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

2 T = Mv2 + UJ-Q-(JJ = M\X — w x p\2 + LJ * Q * LJ. (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 UJ is expressed in
terms of the quasi-velocity x4 and qa by eq. (2.16.5). We obtain

2 T = M\2 ~ 2 Mx * ( w x p ) + w - [ 0 + M (E p • p - pp)\ • u. (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 X1 1 X2 1 X3
an
d these terms remain after setting x1? x2, X3 equal to zero. The third term is the
homogeneous quadratic form in x 4 and qa with coefficients depending upon the
Gaussian coordinates ql,q2. Therefore, we take

2 T = -2 Mx - ( w x p ) + 2 T*,
2T* = <j* • [0 + M (Ep • p — pp)] • u.
(8.3.3)

To obtain the generalised forces we write down the expression for the ele-
mentary work

6’W = F ST + m° • 6

and insert the expressions for the virtual displacement <5r of the pole O and 0 due to
eqs. (2.16.14) and (2.16.18), 6TT vanishing because of the non-holonomic constraints.
We obtain

6'W = (m° - p x F) • (m<$7r4 + 1 a6qa).

The quantity m° — 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 - m , P 5 = M - 1 I , P 6 = M - 1 2 , M - m° - 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

q1 = X5, q2 = Xe. $ = X4 + VW\ (x5 *


k 2
- Xek*1) ■
dT dT
9XrXt ~ d^{knX5 +

dr
si(13X5+
hi £

( 9T ,
II
*

dT ( dr,
9xrXt ■ fe*21+

dT
dX2

dT dT dT*
k\2 + fcl3 ) X4 +765^-X6>
dX 5X3
2

dT , ar, 5T \ 4 dT*
+ +
Wsk2V X4+756
Wtx‘-

dT dT dT
a----kal + a--------&a2 + ~—fca3 = —M (w x p) • ka = M (w x ka) • p.
9xi 5x2 5x3

However, by virtue of eqs. (2.16.18) and (2.16.19), w x ka = mx4 x (u> x pj + x (m x pQ) =

-x4Pa + Q mp ■ h
0
a

— ~X4Pa + a. 1 (8.3.7)
r—r m (b(3i(l2a c*) •
dT „ _ . Q Me = Mp ■ pa<fXi + \aq0 {bpia2a ~ bp2dh
dXr Xt = /i r-<
vM

dT Xt = = Mp ■ PiXi -
Me .p
{b/3ld2l — bp2dii) -
dXr V\a\Xiq
b
dr* 611622 — 12 A2
dx4 VW\ Q>

dT Xt = = Mp ■ P2X4 - Ms .0 (b010*22 — bp 20*12) +


dXr ^iX4?
dT* bnb22 — 612 -i
dXi \/M Q•

Here

6: = m • p

is equal to the distance from the centre of inertia O to the plane on which the body
rolls. Additionally

p paqa = PP>

P'Pa

where p denotes the absolute value of the radius vector p with the origin at point O
and the end at the contact point.
The final form of the equations of motion obtained by Voronets is as follows

d_ dT Me
— MClpp 0?1)2 {bnai2 -b12an) +
dt dQ* 7R
q'q2 (bna22 b220’ll) + {Q2) (p21^22 — b22a\2) Pi,
d dT dT tr\2 dP M E - (, , , . 12 n) + , ,,
JtW " W + PW ~ 7R [q 11012 ” °

q (bi2ai2-b22ail)}-—-------------------p5, (8.3.8)

d dT dT+ P dp Ms rbl2ai2).,. .
7tW~W dqt~W\ 2 ibna22blll>22
~ 2+
2

q (b12a22 - b22a12)\ + ^- ~.%a f = P6.


2
OTi Vl l /

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 % = 0. Voronets
dqa , dqa
771i Pal
~fo ka 1-3—
da
771 dqa Kefir-
2 Pa2
lh da
7713 dqa

CO
Pa3
lb
0
0 i
cos d sin d 0
— sin d cos d 0

(8.3.10)

whilst the rotation matrix making the trihedron Oxyz coincidental with the half-fixed
trihedron is given by

(8.3.11)

Then the matrix

a = (3(31 = (8.3.12)
cos
dqa dqa
(Pal $ ~ kai sin d) —— da (pal sind -f kai cosd) (,pa2 sin d da 7711

(.Pa2 cos $ “ ka2 sin ??) dqa


dqa -b ka2 cos d) (,Pas sin # + ka3 da 7712
(pa3 cos d - ka3 sin d) — dqa
da 7713
cos d)
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, Vo, zo of the pole O in the
fixed coordinate system Oxyz. To this end, it is sufficient to project the vectorial
relationship

YM = ro + P

on the above axes. Here TM = OM denotes the position vector of the contact point and
its projections on axes Oxyz are x, y, 0. We arrive at the formulae

xo = x - (anx' + a iy' + a zf), 2 31 '


yo — y - {OLi2x' + a y' + a^z'), 22
(8.3.13)


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

d dT dt
6 3
ZfT
+ EI>L^-xt = Ps (s = 1 , 2 , 3 ) , (8.3.14)
dxs t=4 r=1 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 s = 1

^ (<■*> X p) 1 + [m 2 (u> x p)3 - TOs (u> X p)2] Xi +


p
[l 1 2 (u> X p)3 - l 1 3 (o> x p)2] x5 + [hv (W X p)3 - l 23 (u> x p)2] Xe = --j^-
(8.3.15)

Here m s ,li s ,l 2 s , (a; x p)s denote the corresponding projections of vectors m, li ,1 2, Co? x p
on axes Ox'y'z'. The time-derivatives of quantities (a; x p)8 are projections of vector (w
x p)*, i.e. the derivatives of the vector u;xp with respect to axes fixed in the body.
Thus, relationship (15) is the result of projecting the vectorial equality

(w x p)* + (mfi + laqa) x (u> x p) =

on axis Ox'. The first factor in the second term equals a;, then

UJ xp + u?x p +a; x (a; x p) = — -^P. (8.3.16)


p ( dp dp
(lal 2 a2
\ dq dq\
p / f dp T d
P\
x / R ' ”aldf~a2W)

(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 — mV = T7p7 (8.3.20)
408 8. Other forms of differential equations of motion

can be determined by the covariant components


d
T~ = P • p7 = M (w xp7) • p-n/p7 • k/3 - u> • P7U> • p+UJ2p-
^
M ky ■ pCl + e u)a ea-yS ~ £l<f—7== (a/3ia 2 - a/? a i) -
7 2 7

. vH
-4= (6Qia72 - &a2a7i) + 1/3 • P40) + >
(8.3.21)
Vlla
°q .
where 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\N\, i.e.
under the following condition

a^TyTg < f2N2. (8.3.22)

8.4 The case of a body bounded by a surface of


revolution
It is adopted that axes Ox'y'z', Oz' 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 = u> • [0 + M (Ep • p - pp)\ • u> =A1a>f + Ricdl + C\u\ - M (w • p)2 ,


(8.4.1)
where
Ax = A + Mp2 = A + M (z/2 + v2) (8,4.2)
etc. whereas the scalar product LO • p due to eq. (3.19) and the formulae of
Sec. 2.16 for the surface of revolution is given by

cop — • p + qala • p — Qe + Cp 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 = A\UJ2 + (C — A — Mz'2) L— Mv2u2 - 2Mz'vueus, (8.4.4)


where, by virtue of eq. (2.16.26),
J2 = f)2 + Cp2 sin2 a + k2s2,
us = Q cos a — Cp sin2 a,
ue = (fl + Cp cos a) sin a.

(8.4.5)
8.4 The case of a body bounded by a surface of revolution 409

The potential energy of the weight is


II = Mge — Mg (v 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
d dT njr, ,.
— —— — M iv cos a — z‘ sin a) si 2+
dt oil

M (y sin a + z' sin a) (—kv + sin a) sp — 0,


d dT dT ,.
— —--------- —h M (vcosa — z sina ir-
dt os os ► (8.4.7)
<9T
M {y sin a + zf sin a) ftp sin a — s
^n a
= —Mgk (u cos a — z' sin a),
d dT dT dT
— —-----——h M (is sin a + z' cos a) vsflk + -^rrsk sin a = 0.
dt dtp dtp dil
Here k denotes the meridional curvature. If A = B the coordinate p does
dT
not appear in T and term in the latter equation vanishes.
dp
Instead of the second equation we can consider the energy integral

T + n = E. (8.4.8)

In the case A — B, by using independent variable s, we can cast the remaining


equations in the form
ddT
— —— — M \y cos a — z sina il+ dsdQ
M (y sin a + z' sin a) (—kv + sin a) p = 0,
d dT dT (8.4.9)
— —- + M (v sin a + z' cos a) vflk + —k sin a = 0.
ds dp dft

These two first order differential equations determine quantities p and


ft 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
z' = h + a cos a, v = a sin a,
where h and a denote the distance from the centre of inertia O and to the
geometric centre of the spherical surface and its radius, respectively. Then
k- -1
, s = aa and equations (9) take the form
d dT

— —— + Mahil sin a da dil 0,

d dT , fa \_. dT
01
— —- + Mah T + cos a H2 sin a + da dp \h sum 0. (8.4.10)
/ dil
410 8. Other forms of differential equations of motion

Removing MahQ from these equations we arrive at the relationship


d dT \(a \ d dT dT .1 n [{h + ™Vtedn-d!isma\ =0'
It admits the integral
dT /a \ dT , .
(8.4.11)

which is a linear relationship between Cp and 0 and enables one to express these
quantities in terms of each other. Eliminating now Cp from the first equation in ( 10)
we arrive at the first order differential equation linear in 0 which is always
integrable in quadratures. The quantities Cp and 0 will be expressed in terms of a.
Though the next step for obtaining the dependence 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 = 0. 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 ^ B ^ C can be reduced to
quadratures.
In the latter case

2T = AnQ,2 + 2A12n<p + A22<P2 H- A336?, (8.4.12)

where
An = C cos2 a + A sin2 a,
A12 = —{C — A) cos a sin2 a,
(8.4.13)
A22 = (.A cos2 a + C sin2 a + Ma2) sin2 a, A33 = A +
Ma2
8.4 The case of a body bounded by a surface of revolution 411

and differential equations (10) yield dT

d dT
—— = (3 = const, — —- + Ma2Q sin a 4- /?sina = 0. (8.4.14)
oil da dip
Eliminating Cp from the second equation and rearranging the terms we cast this
equation in the form

an A(c + Ma2) / 0 \
da AC 4- Ma2 (A sin2 a 4-C cos a) \
2
A) (8.4.15)
Integration yields
(3 cos a
A ^ Vl 4- A cos2 a ’
where 7 is a new constant and A is defined by the equality (8.4.16)

C - A Ma2
~ C 4- Ma2 A '
The first equation (14) is written now as follows
/ A\
.1
~ ~ 2 sm
a

(3 cos a
A-C
cos a 4- 7- (8.4.17)
\/l 4~ A (

Let us consider a particular solution

Q = 0, a = ao,

obtained by means of the relationship between the constants /?, 7, E, ao derived from
the energy integral. In this case the trace of the contact point is a parallel circle a = a0.
By means of eq. (2.16.25) we obtain

$ = Q 4- (p cos ao, $ = (Q 4- Cp cos a)t 4- $0

and, due to eq. (2.6.10), the trace of the contact point is the circle

a<p sm ao a(p sm ao
x = XQ 4-
Sl + (p cos ao
sin$, y = yo —- - -:-------cos n
v.
il 4- <p cos ao
Finally, p — (pt+p0. The solution has seven constants ao, /?, 7, $0, ^0,2/o, 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 (<j® = 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 (3 = 7 = 0 and thus Q =
0. 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 $ = 0, i.e. $ = $0 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 <p = 0, $1 = 0 takes the form

2 (E - Mga) _ .
a=± A + Ma2 ~ a°’

an
thus, a = aot + ao. The solution has six constants a0, do, $0 d 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 ^ 0.
Instead of eq. (13) we obtain the following expressions for the coefficients of the
kinetic energy

An = C cos + (A + Mh ) sin a,
2
a 2
'2

A12 = (A — C) cos a + Mh (a + h cos a)] sin a, 2

(8.4.18)
A22 = ^Csin a + A cos a + M (a + /icoso;) ^j sin a,
2 2 2 2

*
Integral (11) is now given by

[Aa sin2 a + C (h + a cos a) cos a] Q +


[Aacosa — C (h + a cos a)] (p sin2 a = /^/i, (8.4.19)

and the first equation (10) takes the form

-7- (Antt + Ai2(f) + MahQ sin a = 0. (8.4.20)


aa

The calculation simplifies if, instead of Cp sin2 a:, we enter the projection uos of
the angular velocity vector LO of the body on symmetry axis Oz' due to eq. (5)

UJS = cos a — Cp sin2 a. (8.4.21)

Equations (19) and (20) are transformed to the form

AaQ — [Aa'y — C (ft + ay)] UJS = Pih, (8.4.22)


8.4 The case of a body bounded by a surface of revolution 413

— {[A + Mh (h + (27)] Q — [(A — C) 7 + Mh (a + /17)] ^3} = MahQ, <27

(8.4.23)

where cos a = 7.
Rewriting these equations in the form

Aa~r~ — [Aa'y — C (h + a7)] - (A - C) auo3 = 0,


a7 a7
[A + Mh (h + a7)] ^ — [{A — C) 7 + Mh (a + /17)]
<27 <27

(A — C + m/i2) CJ3 = 0

and removing —- we obtain the following first order linear equation


<27

d'j
AC + Mo2 A (1 - 72) + MC (h + a7)2
+

[CMa (h + a7) — AMa27] CJ3 = 0. (8.4.24)

We easily find that

-1/2
^3 = 02 AC + Ma2A (1 - 72) + MC (h + a7)5 (8.4.25)

Now we obtain from eq. (22)

n=M + 02 [Aa') — C(h + 07)]


77^- (8-4-26)
Aa
Aa AC + Ma2A (1 - 72) + MC (h + 07)2

The expression for the kinetic energy in terms of ft, cus and a is transformed to the
form

2T= T^{[ A + M (ft + o7)s ft2 + (8.4.27)

^22 - 2 [^7 + M (h + aj) (a + hj)] flw3 + ^43372


1 —7
Obtaining the energy integral and inserting expressions (25) and (26) for the quasi-
velocities $1 and into it we reduce the problem to determining
7 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

7=!-^=!-y(1-^|+...,
(8.4.28)
[Aa
— C (a + h)}2 M C-A
1____1 A

(N ' +
AC 414 8. Other forms of differential equations of motion
+
A + M(a + h)2 [I {C + M(a + h)uf 4 1
77 being small.
a2 + Denoting the values of a;3 and ft for 7 = 1 by r0 we obtain
4 A + M(a + h)2 4 y
^ AMa2 — MC (a + h)a
^3 = r0 V + -- ,
C A + M{a + hY
► (8.4.29)

ft = r0 < 1 - V + ---

with knowledge of the higher order terms not being needed.


Calculating the expression in braces in formula (27) up to the terms of
Tj2 we have

2T (1 - 72) = \A + M (a + hf 1 i)2 + 2Crfa + r2 {£ + M(a + h) o}2 2


L J
A + M(a + h)
(8.4.30)
Returning to angle a we obtain

„2'
2Taz | 1 - 2L ] = 2-2 ,
A + M (a + hy aa+

Cr\a2 ^1 o
[<? + M (a + h) a}2 a4 A +
M(a + h)2 4
or

2T = [A + M (a + h)2 .21 2 J M (& d- ^) &]


^^ + c^+cr02.
^ + M(a + /i) 2

(8.4.31)

The potential energy is equal to

II = Mg (a + h cos a), (8.4.32)

so that the energy integral is written in the form

a2 = const.

(8.4.33)

Taking the time-derivative yields

n2
C ~h M (CL ~h /i) U 1 Crl-AMgh ±A +
" +<f + a = 0. (8.4.34)
4+M (a + hf M{a + hf
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

1/2
C + M(a + h)a\2 1 Cr%-4Mgh
k= (8.4.35)
A + M(a + h)2 4 A + M(a + h) 2

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 g q , . . . , g n . Equations for the non- holonomic constraints, if
they exist, are written in the form
n
aksQs + afc = 0 ( f c = l , ( 8 . 5 . 1 )
S —1

where ctks and ak are functions of the generalised coordinates and time. Equations (1)
should be resolved for l generalised velocities gq, that
is the deficiency of l x n matrix (7.10.5) should be zero. Then expressing l generalised
velocities in terms of the remaining n — l ones we obtain the system of equalities

qr — frr,z+ig,i+i T • • • T br<>nqn + br (v — 1 , . . . , / ) . (8.5.2)

Equations relating variations of the generalised coordinates


n

Y,aksSqs=0 ( f c = l , . . . , Z ) , (8.5.3)
S= 1

allow the variations <5gq,..., 6qi to be expressed in terms of n — l variations 6qi+i,. . . ,


Sqn, the latter variations being independent. By analogy to ( 2) we obtain

n—l
6qr = br,i+s$qi+s (r = l,... ,1). (8.5.4)
S= 1

Differentiating eq. (2) we find expressions for the generalised accelerations

n—l n—l

Qr ~ ^ ^ br,l+sQl+s + ^ ^ br,l-\-sQl-\-s T W. (8.5.5)


S=1 S = 1
416 8. Other forms of differential equations of motion

These equalities will be used to eliminate qr from expression (1.3.9) for the
acceleration vector of particle Mj. We have

w dr i drj
i= ~Q m • •
dqi+s Qi+s + • • • ,
(8.5.6)
m= 1 dqr

where dots stand for the components which do not depend on the generalised
accelerations. Making use of relationships (5) we recast the latter equality in the form

l 0 n—l n—l
OYi dr i
Wi
-++

n l / £V l r\ \

Denoting

di + ^2 ,1+8 = C< ,1+s (i = l , . . . , N ; s = l , . . . , n - l ) (8.5.7)


dqi+s dq,

we obtain
n—l
W i = ^2 °i,i+s<ii+s + ■■■ (8.5.8)

and since the unwritten terms do not depend on the generalised accelerations we
arrive at the equalities

<9wi .......
o** — Ci,l+s (8.5.9)
dqi+s

which are used in what follows. It is easy to understand that the virtual dis -
placements Sr j are expressed in terms of the independent variables 6qi+s by means of
the linear relationships with the same coefficients Citi+S. Indeed, using (4) we have

drj
Sri Sqi+s
dqi+s

n—l dqi+s ~{dqr r,Z+S


£ Sqi+s

or, by eqs. (7) and (9),

n—l n—l
^ ^ci,i+sfiqi+s — ^ ^ fiqi+s-
s=1 ^l+s (8.5.10)
s=l
8.5 Appell’s differential equations 417

Applying relationship (10) to rearrange the left hand side of the fundamental
equations of dynamics (6.3.2), we obtain

N N n—l
dwi
Y rriiWi •8vi = Y &Y^qi+s
2=1 2=1
n-l N
dwj
E * « + « E rriiWi (8.5.11)
diji+s
S= 1 2=1

Noticing that

N
E <9w id1
N
dS (8.5.12)
rn1vsl • ‘TJT;— = ——- ^ rnlwl • w*
. -i @Qi+s &Qi+s 2 dqi+s
2=1 2— 1

we arrive at the equality


N n-l dS
E m*Wi • = E Qql+ (8.5.13)
2=1

in which 5, 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 q\ , . . . , qi 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
Fi
6'W = EFi-^i = E ‘E *,«+»*«+» = E Qi+‘ <b+'- (8- 6

5.14)
2 =1 2=1 S= 1 S = 1

Here
N
Qi+s = E Ci,i+s ■ Fi (s = 1 , . . . ,n — l) (8.5.15)
2 =1
denote the generalised forces corresponding to the independent variations of the
generalised coordinates. According to eq. (7) they can be recast as follows

N +'Ejrb
dTj
Qi+s — ^ fridQr
dqi+s


or, due to eq. (5.1.3),
i
Ql-\-s = Ql-\-s 4“ ^ ^ br,l-\-sQr
r=l
1, . • • 5 Tl /) ,
(8.5.16)
418 8. Other forms of differential equations of motion

where Qz+s, Qr denote the generalised forces calculated under the assumption that all
variations are independent.
The fundamental equation of dynamics is now brought to the form

(8.5.17)

Since the variations 6qi+s are independent we obtain Appell’s equations of motion

dq i + , ~ ® l + a 0 (8.5.18)

first suggested in 1899, [2]. Their number is equal to the number of degrees of
freedom. They are second order differential equations for the generalised coordinates
qi+i , . . . ,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 equations of the
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 instead of the
generalised accelerations.
Let ui,...,wn denote quasi-velocities, equations for the non-holonomic constraints
being cast in the form

= 0 (s = 1, . . . , / ) . (8.6.1)
The velocity vector of a system particle due to (1.5.19) is given by
8.6 Appell’s equations in terms of quasi-velocities 419

Taking the time-derivative we obtain the acceleration vector

n l
~ ftr
Wi = Y] ^— —Ui+s + ■ ■ ■ , (8.6.3)
dni

dots standing for terms independent of the generalised accelerations. The
consequence of eq. (3) is the following relationship

<9w i dr i
. • (8.6.4)
diui+s dir i+s
On the other hand, taking into account that, in the case of non-holonomic constraints
(1),

STTS = 0 (5 = 1 , . . . , Z ) , (8.6.5)

we obtain due to eq. (1.6.14)

n—l 0 n—l o
Sr, = £ -p6*l+, = E P-6*l+.. (8.6.6)

Thus
N
rriiWi • <Sr* = ^ N ' ^2 ft' 1 SlTl+s = ^2 ^l+s 5 2 miW'
i= 1 8 = 1 OLJl + S 8=1 i=l
du l-\-s
i=1
n—l Q i AT

L S
s=l ^ i=l

or
N n—l
dS
T rriiWi -6ri = y —------------------6ni+s, (8.6.7)
2=1

S being the energy of the accelerations. By virtue of eq. (5.1.6) the elementary work
of the prescribed forces can be represented by the equality which is, due to eq. ( 5),
written in the form

n—l
6’W = yPl+s6irl+s, (8.6.8)
S = 1

where Pi+S denote the generalised forces corresponding to the quasi-coordinates.


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

6*1+8 = 0. (8.6.9)
420 8. Other forms of differential equations of motion

By virtue of the independence of variations <$ 7 n+s we obtain


Appell’s equations in terms of quasi-velocities

dS
dui+s ■l+s (s = 1 ,n, —
. l). (8.6.10)

This system of n — l first order equations contain the first


time-derivatives of quasi-velocities cj/+s, the quasi-velocities
and all generalised coordinates qn and need to be completed
by n first order equations (1.5.25)
n—l+1
Qr = ^2 br,i+kUi+k (wn+i = 1 ), (r = 1 , . . . , n) ,

(8 .6 .1 1 )
k=i

determining the first time-derivatives of the generalised


coordinates in terms of the quasi-velocities.
Clearly, Appell’s equations (1 0 ) are applicable to
constructing the equations of motion for holonomic systems.
As will be shown below, Appell’s equations (1 0 ) 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 constraints which complicates the
structure of the expression and the very construction of
equations requires special attention to indices. While dealing
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 systems, 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 application of the second rule
requires knowledge of the three-index symbols and an
expression for T obtained under removed constraints, while
the application 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 of the second
kind (with non-holonomic constraints being absent). The rules
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
accelerations. With non-holonomic constraints (5.1) being
absent, Appell’s equations (5.18) according to (5.1) take the
form
dS
g^-=Qr (r= 1 , . . . ,n).

Using eq. (4.10.6) we obtain

dS = n n 71

TtfT Asrqs + EE [s, k\ r] qsqk +


s= 1 s= 1 k=1

dAsr dB, dBs dBr


E qs + dt (8.7.2)
dt +
dqt dqr dqr
S=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
= £s(T)=Qr (r = l , . . . , n ) .
(8.7.3)
Oqr
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
where dots designate the terms which do not contain the
generalised accelerations. Energy of the accelerations is now
dependent upon the generalised accelerations #/+&,
appearing in eq. (4.10.6) explicitly as well as appearing in S
in terms of qr. For this reason
dS ( dS \ dS dqr
dqi+k \dqi+kj dq rdqi+k

or
l
dS
dqi+ £l+k (T) + brj+k£r (8.7.5)
k (T)
r=l
ail ai 2 .. • an
^21 a22 • • • a2i

an ai 2 .• • an

(8.7.8)

and the determinant la*! = |a'J. The algebraic adjunct aqr of


element aqr of the q — th row and the r — th column of this
determinant is equal to the algebraic adjunct aqr of element aqr
of the q — th row and the r — th column of the determinant |a'J
of the transpose matrix a*. The solution of the system of
equations (8 ) has the form
™ ^ ^ ^ ^ ^ Q<qr
Qr — ~ ^ ^ Ql+k } ^ I ?aq &q,l+k ~ ^ ^ aq~ Ti (8.7.9)
fc=l q=l |a*' 9 = 1 *'
|a

thus

^ Qqr
br,l+k ~ ^^1
7^q,l+k->
(8.7.10)
\a*
q=1 1 1
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 qi+k- This is implied in the
form (5.18) of Appell’s equations which is written now as
1— = Ql+k (k=l,...,l), (8.7.11)
oqi+k
Sdenoting the expression for the energy of accelerations after removal of
quantities qr.
Let T denote the result of eliminating the generalised
velocities qr, given by eq. (5.2), from the expression for the
kinetic energy. However it would be a grave error to assume
that under non-holonomic constraints

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 qi, . . . , qi do not appear in the
expressions for T and coefficients brj+k- Then we have
T(qi+i,... ,qn;qi+i,... An) = T(qi+1,... ,qn;qi,... A1A1+!,••• An)
and

Now accounting for that


424 8. Other forms of differential equations of motion

we obtain
i
£l+k = £i+k (T) + ^ brj+k£r (T) +
r=1
(i dbrj+m ^
/ O - I Or,l+k x. o (8.7.13)
#i+n
^ dqr ^ ^ dql+k
But
TL l 0 7
l _
r,l+k
<?0 r,z+ra .
~ h d(ii+k qi+m’
so that due to eqs. (6 ) and (5.15) we
obtain
dbr,l+m
(7i+n
M ) - ««+£§£fr;( , <»'«+.
f ?w1 %+fc
(fc = 1 , . . . , n — Z) (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
_ ^ 8T +1 fdbr,i+k dbby following
r,l+m\ .
correction terms
l+k
hi dklr hi \ dqi,+m dqi+k ) qi+n (k = 1 , . . . , n — l) .

(8.7.15)
These terms vanish
when
^^r,l+k —0 (r = 1 , . . . , Z; m, k — 1,... ,n

^^r,l+m — l) .
dqi+ m dqi+k ictions fr (qi+\, . . . qn) such tha
Or,l+k
But in this case —
dqi+k
there exist functions fr (qi+i,.. ■ qn) such
and the constraint equations(r =(5.2)
1 , . . . will
, Z ;be
k =integrable
1 , . . . , n —(as
l) the
constraints were assumed to be stationary) since
. dfr . dfr . •
Qr = ~--------qi + 1 + . • • + -^-Qn = fr
dqi+k uQn
or
qr = fr (qi+i, ■■■qn) ( r = l , .
The coordinates qi,... qi 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 slipping is, due to eq. (4.13.3), given by
7 (i2 i2\ 2 :21
5= - I dq + d) 2 1 + -
d; 3 ,
where we have omitted the terms corresponding to the
equations of the non-holonomic constraints, i.e.
CO4 — 00 5 = 0.
The left hand sides of Appell’s equations (6 .1 0 ) are equal to
dS 7 ^ 2 2■ dS 7n, 2 : dS 7 „ 2 : ,
—— = -Ma CJI, —— = -Ma2 CJ2, —— = -ALa2 uo3 .
/on N

(8.8.1)
1 UJI 0 Co2 0 Co3
The generalised forces are determined by expressions (2.4).
Finally we arrive at equations of motion (2.5).

8.8.2 Ring
The energy of accelerations is given by formula (4.13.12)
which yields
dS
duji Ma2 -Cui -F 2CJ2<^3 — -<^2
cot d
dS
dCu Ma2 T. 1 (8.8.2)
-002 - 0J30J1 + -CJiCJ2
2
gTj- = Ma2 (2CO3 - 0J1UJ2).

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
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 constraints2 have the form
T= (±0 + yl) + "Ma ($2 + + (p + 2<^cosi?^ ,
2
(8.8.3)

(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 XQ and yo from T by means of the constraint
equations we obtain

Adopting the following numbering of the generalised coordinates


Qi = zo, Q2 = 2 /0 , Q3 = Q4 = Qb =

we obtain

The generalised forces Q2+r given by eq. (5.16) are equal to

Now we need the values


r =1 r =2 r =3
—a ($cos ip+ a (ip sin $ sin 'ip
s =1 V
T a cos 'ip (ip + p cos
p sin d sin 'ip) —i) cos $ cos ^
—a (iOsiw'ip— —a (ip sin $ cos 'ip
s =2 a sin ^p (ip + p cos -f $ cos $ sin ip^j
Y
V
p sin cos 'ip)
Qi = Mxo - Vu Q\ = My0 - V2, Q* = 0, >
—M (xo sin ip — y0 cos 'ip) + (V\ sin ip — V2 cos ^
Qt = a >>
M (XQ cos 'ip + yo sin ^p) — (V\ cos 'ip + V2 sin ^ .
Ql = a >

Since
— = Mx o, — = My0,
dqi dq2
we obtain the following values of the double summations on
the right hand sides of equations (7.14)

r=1
r=2 —Ma2p (ip -\- p cos ^ sin
r=3 Ma2i) (ip + pcosfi'j sin$.

After simplification Chaplygin’s equations take the form


:Ma2 ip cos — (pi) cos = rhs
-Ma2 + (pip sin ^ = TTIN — a (V\ sin 'ip — V cos % 2

► (8.8.5)
22- 7.
\
ip sin2 $+ -ip + -Up sin cos $- - -(pi) sin $ )
( 5 5 5 J
= ms — a (Vi cos 'ip + V2 sin
Clearly, equations (2.9) take the same form when xo 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

The complexity of the equations of motion (8.5) and


calculation for such
a simple problem as the rolling of a sphere explains why
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 — OJ\f and
the double area a of the sector OMQM swept over by the
position vector measured from the initial position OMQ.
(8.8.7)
so that the kinetic energy of a particle of unit mass is

(8.8.8)

Relationship (7) is a non-integrable equation defining the


quasi-velocity a in terms of the generalised velocity Cp.
Designating

(8.8.9)
The elementary work of the force applied to the particle is

Now we construct the equations of motions by means of (1.13)

d dT dT
“77 -2- +712^- -^1 = ~
dt 2 uuj2
duo rr
Calculation yields
2d2 a2
a 2rb 2or _ 1
o TT o o TQ rn
8.8 Applications to non-holonomic systems 429

Hence we obtain the following equations of motion

(8.8.10)

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

wr = r — rCp — r — —, = - yr pj =

we obtain expression for the energy of accelerations

and Appell’s equations (6 .1 0 ) yield the above equalities (1 0 ).

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 (p of shaft 2 (due to acceleration or
2

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 (p is restored as a result. The differential
2

equations of motion of the system are required.


The position of the system can be defined by three
generalised coordinates, namely the rotation angles cp of 2

shafts 1 and 2 and the distance x of the coupling D from joint


L. The distance p — SQ is expressed in terms of x. This relation
0\D = OxL - x, 02K = 02S + p-KQ

and, as 0IL,02S,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
f*Pi = 2- (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

2 2 (®B~2V2 + mBp , -mDx2.


2
(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
cos a.
x = 21
Let us consider the projections of acceleration w^v of the
sphere centre on three orthogonal directions which are the
directions of the rod LN, perpendicular 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,
la — 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
± 0 / # 0 # 0 ' o o \
—2TYINI (d — 2aCp2 sin a cos a + ip2 sin a + 4ip2(p2a s^n a cos a) •
(8.8.14)
The generalised acceleration ip is removed from the 1

expression for the energy of accelerations by means of the


equation for the non-holonomic constraint ( 1 2 ), to give
R\^ R PP 2
Pl = P P2
P
R /\.9 'P
2

p\ ~ ^2 ..2J. \+ • • •
2-P2P

The justification of this procedure was discussed in Sec. 8.7. We have


x x* cos a
'+4■Z2 sr
a= — , OL = ~
21 sin 21 sin a sin3
a
x a xx cos a
a2 =
4/ sin22 a+ 41.,0
2 3
sin
. A4 ---------1 " . . .
a
Now adding eqs. (13) and (14) we obtain S in
the form
&AT—-T2 + ©C + ©B-y + 7} N (41 m 2
- X2)
5
= 2^
(x-c)
1
QAH m 1. 2ui TV l2
VVX[ 2 2 ^ _ c)3 + NX j + 2* \mB +mD + 4p _ ) + X 2

2l2x2
mjsfxx T --------
~(P2 9
• 2 (8.8.15)
22 (4Z2 —
We determine now the generalised forces. Let (p% and XQ
denote respectively the values of Cp and x under a stationary 2

rotation. In this case the spring of the governor has length h + 61


greater than the natural one, whereas the spring of wheel B
has length Z2 — ^ 2 less than the natural one. Under a deviation
from the stationary rotation the lengths of the springs become
Zi + Si + x — £0, ^ 2 ~ $2 T P ~~ Po — ^ 2 — $2 T x — XQ.
The changes in length, referring to the natural states, are Si +
x — Xo and
S2 — x + xo, respectively, so the potential energy of the elastic
forces up to
a constantn term is
2 2 1 2
1 9
= (°1 + C 2 ) ( x - x 0 ) + ( c i 8 i - c 2 6 2 ) ( x - x0) • (8.8.16)
432 8. Other forms of differential equations of motion

The elementary work of the moments mi and m2 applied to


shafts 1 and 2 , respectively, is due to (1 2 ) given by
8rW = mi8(p + m 8ip
f mi —----------
1 - -b 2 ) 8cp
2 2
= m
2

V x~c )
and the generalised forces corresponding to the coordinates
(px and ip are 2

Qtp, =
ll
+ rn2,
mi x—c
Qx = - (ci + c2) (x - x0) - (ci8i - C 2 8 2 ) .
The differential equations of motion obtained by means of
scheme (5.18) have the form
R
0 (x) ip ~ $ (x) X(p = mi-----------b m2 ,
2 2
x—c
1 2 l2x (8.8.17)
m(x)x+ -mNxCp\ + —---------------mNx2+
* (4l2 - x2)
(Cl + c2) (x - x0) = -Ci^i + C 2 6 2 , J

where it is denoted that


0 (a;) = QA— ^ — 2 + ©c + ©s~2 + (412 - x2),
(x — c) al
, x. @AR 2

v ) = 7-----7 3 + mNX, m (x)


f
= TUB + mo +
x 2m^l2 > (8.8.18)
(x-c) 4/ 2 — x 2

In order to prove the result obtained we construct the


equations of motion by Chaplygin’s approach. This is allowed
since (px enters neither the expression for the kinetic energy
nor the equation for the non-holonomic constraint. The kinetic
energy of the system is
[0 (x) p>2 T 777- (x) X2]

where the first term


R2
0A
(x-c)
of the expression for the reduced moment of inertia 0 (x) is
obtained by eliminating Cpx by means of equation (1 2 ) for the
non-holonomic constraint from the term

Chaplygin’s equations are as follows


£<P2 (^) — Qtp + ^2 ’ £x (r'j = Qx + Rx
2
(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 -ft

R2

RV2 = OAVXX— ,sX(P2’>
dx V x — c (x - c)
« R
-2
(8.8.20)
d R
R.r -0 A¥>I¥>2 -T: (x-c)
dx V 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 = 0 , x — 0 , = ^25 ^2 = 0.
<p2
We arrive at the relationships
R 1 2
m?-------hwo = 0, -rayvico^o + c\8\ — c2<$2 = 0.
-----------(8.8.21)
XQ —c 2

Entering the quantities which define the deviations from the


stationary
regime x0 <P2

assuming them to be small and accounting for ( 2 1 ) we obtain


the system of linear differential equations
„,N• _ , . . m^Rxo m\R
0 (x0) n - X <$> {xo) q + —TTQ-Q = -TTo (Mi - M2 ),
0
MO^ 2 P&2 (8.8.22)
m (x0) q + -mN(p% (q + 2 1 2 ) + (ci + c2) q = 0 ,

where
m i— m2 — m 2
Mi = m o
M2 — m;0
Let us consider the corresponding homogeneous system Mi
=
M2 = 0. Looking for the particular solution in the form
17 = Q0eXt, q = q0ext,

we obtain the characteristic equation for A

0 (XQ) ( C i + C 2 + —rriNtp2 ) +
Q(xo)m(x0) A3 +

(#o) <£2 A + rriN—m2(/?2 = 0.


(8.8.23)
Mo
434 8. Other forms of differential equations of motion

Here is replaced by means of the first equation of the relative


equilibrium (2 1 ). A dash-pot with a resisting force proportional
to the coupling velocity introduces the term with A 2 into eq.
(23). Then the second equation in (2 2 ) gains a term with q.
Stability of the stationary rotation can be achieved when
power > 0 is supplied to shaft 2 whereas shaft 1 acts
as its
receiver, i.e. < 0. The rotation is stable if the coefficients of
eq. (23)

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 coordinates. Using expression (4.1.6)
for the kineticnenergy we find n nn
dT YAUPk, d dT
A*k“>k+ Y! dKk
dus fc= 1 dt dus d'Kr LUr(^k •

Substituting this into eq. (1 .8 ) yields


nnn
A
Y *sk^k + YJ Y! YJ 'YtsArku;kUt +
k=l k=1t=1
r=1
n n2 ^ 2 ^ UrUJk — / dA*k 1 9A*rk = Ps (s = 1 , . . . ,n)
~ d'Kr 2 dnr
r= 1 k=1
Applying notation (4.10.9) we can write the double sum in
the form

(dKk ldKk

h h “2 *Tr
dA kr
d'Ks ^r^k •
r=l k=1 +
dwr r=l k= 1
Then, we
obtain
YlA*sk^k ^lrtsA*rk + \t,k;s]v l = Ps (s = 1,... ,n).
k= 1 k=1 t=l (r=l J
■ (8-
9.1)
Obviously we would arrive at the same equations by
constructing Appell’s equations (6 .1 0 ) with the help of the
energy of acceleration (4 .1 0 .1 2 ). Given l equations of the non-
holonomic constraints
Wm — 0 (m= 1 , . . . , / ) , (8.9.2)
8.9 Explicit forms of the Euler-Lagrange equations 435

equation (1) takes the form

53 ^sk^k + 53 53 | 53 ^ts^-rk + [^5


U kU t s
]7r [ _
fe=Z+l fc=Z+l t=l+l V r=\ )

(s = Z + l , . . . ,ra).

(8.9.3)
The relationships
n n n (n

53 + 53 53 153 ^ ^*fc + ^
UkUt 7 m
fe=Z+l fc=Z + l t=l+l lr=l
= Pm+Am (m = l,...,Z),

(8.9.4)
in which tok should be replaced by their expressions from eq.
(3) serve to
determine of the generalised constraint forces A m.
As an example let us determine the generalised constraint
*34 -Ma,> A^15 -Ma sin$.
The three-index symbols of interest are 7 2 4 and 7 2 5 which are,
due to eqs. (2.10.28) and (2.10.29), given by
724 1 725 1
sin $ ’ lzo
sin d
Analysis of formulae (4.10.9) indicates that the only non-zero
bracket is

==
Then we obtain -Macos^

A4 = Ma (—CJ3 -f- LU1UJ2),


W 2^3 9 (8.9.5)
\
( -u 1 sin $---:—- -
uf cos V .
The same expressions can be obtained by applying the
theorem on momentum. Using the ”half-fixed” axes described
in Sec. 2.10, we have
Q = Mv0 = M (cj4n + ^5ni -f ii3).
Taking into account
436 8. Other forms of differential equations

of motion
we find
Q=M ^2^5 UJ2^4
sin$ \ ni + zi
3

sini?
or, by eq. (2 .1 1 ),

Q = Ma
(—W3 + LJ1W2) n — (clq sin $ + UJ\ cos
^2 ^$3 +\ ni + -13 .
z.
sin$ / a

The projections of the resultant vector of the external forces


V on directions n and ni are the required forces.
The derivatives ^ 3 and UJI in eq. (5) should be replaced by
their expression using equations of motion (2.14). Then we
have
-Ma '
1
A
4
4
-UJ2 + UJ \
sin^ - sin$ > (8.9.6)
COS $ + CJ2^3
A5
3 g
— - — sin $ cos d
4 a

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 (ylx + VQ + ^0 3 ) + 2M [(^0 2 ^ 3 ~ ^0 3 ^2 ) xfc +


2

(V03V1 - ^0 1 ^3 ) y'c +
(^0 1 ^ 2 - ^0 2 ^1 ) z’c] + ©i^i + ©2 ^ 2 + ©3 ^3 }.
(8.10.1)
Here ^0 1 ,^02,^03 and 002,^3 denote projections of the velocity of
the pole O and the angular velocity on principal axes of inertia
Ox'y'z', respectively, and x'c,y'c,z'c are the coordinates of the
centre of inertia referring to these axes. The necessary three-
index symbols are given by eqs.
(2.10.3) and (2.10.8), the quasi-velocities ^ 01,^02,^03 having
numbers 4,5,6 by virtue of (2.10.4). Now we have
dT
dv01 M (v 0 i + z'cu2 ~ Vc^3),
dT ► (8.10.2)
did 1
M
{y'cv
03 - 4^ 02) + © J a ; i

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
731 — ~ 721 — 761 724
751 — 1? 734 - — 1?

Turning now to the Euler-Lagrange equations we


obtain d dT dT dT
dt dv0i + ^2 —------^3 T; = Vi,
dv03 dv02 (8.10.3)
d dT dT dT + dTvos dT
j t d ^ ~ +V
d^ ~ d^r = mo 02
'02

and other analogous equations. The right hand sides are


written in accordance 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
Voi + U2V03 - ^3^02 + ^2z'c - ^3y’c - (w2 + ^3) +
Vi
“>i (u2Vc + ^3 z'c) = —, (8.10.4)
the other two being obtained using a cyclic permutation of
indices. The
first equation of the second set is
©l^l + (0g — ©2) CO2^3 T
M [y'c (V03 + ^1^02 — ^2^03) — z'c (T02 + ^3^01 — ^1^03)] = TTI°.
(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
2
'C X (v +Cc? X VQ^ H~ “Cc? • 0^ • Cc? + (JJ • ((A? X 0^ • Cc?) .
M CJ

Using the differentiation rules
_d d (aibi + a b + a 6 ) = b ,
_ ab _ 3 3 s 2 2

dat das (8.10.7)


—a
d" x b = i i x b = i352 - i2&3,
oas
and taking derivatives with respect to quasi-accelerations Vos
and cbs, we obtain Appell’s equations in the form
OS
dvos
= M {v0s + (co x v0)s + (d? x r'c)s + [u x (u x r'c)J} = V8,
dS X VQ^ + (0^ • UJ + Co X 0^ • Cc?) t
dujo =M =mo
(S = 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 VQ x VQ + x r'c + x (u? (8.10.8)

0° • Lj -f- Lj x 0° • uj H- Mr'c x ^VQ ~\~OJ x vo^ — m^.


(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
v0 +U> xv0 + w x r ^ + w x ( c / \
Q=M

that is eq. (8) expresses the theorem on momentum.


The theorem on change in the moment of momentum has
the form
K° = m0,
where the moment of momentum about fixed point O is given
by expression
(4.8.10) . Taking time-derivative we obtain
K° =r0x Q - | - v 0 x Q + M ( c t > x Y'c) x v0 +
Mr'c x ^$o x vo) + @° • CJ + x 0° • = m^,
(8.10.11)
where the rule of differentiation of tensor (4.3.13) and
relationship (4.3.12) are applied, to give
(0° x CJ) • UJ — 0° • (u> x OJ) = 0.
The known theorem of statics and eq. (10) yields
= m° + ro x V = m° + r0 x Q.
(8.10.12)
On the other hand
vo x Q + M (u> x r^) x vo = vo x M ^v0
+CJ x + M (co x r'c) 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 momentum and angular
8.10 Equations of motion of a free rigid body 439

These are differentiation of vectors and tensor 0° 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 O is made
coincident with the centre of inertia C. Then r'c = 0 and the
equations of motion of the free rigid body takes the form
M +u> x vcj = V, 0C - u> + u> x 0C • u = mc. (8.10.13)
Let us mentally separate a part Si of the body S using
surface a. Let Mi denote the mass of this part, T' = CC\ and r'K = C

Cll the position vectors of its centre of inertia and the point on
surface <r, respectively, and 0^1 the inertia tensor of Si 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 S at
points on the body S' by V(i) and m^, respectively, whereas
those of the reaction forces acting on surface a by R and MK.
Mi v(j -\~LO x V(7 T UJ x T (o x ('LO ) V^i) T R..X — (8.10.14)

Taking into account eq. (13) we obtain


R
= -v(1) + + M,[(ix rg + ux (w x rg)] .
(8.10.15)
By means of theorem (12) from statics
mg + MCl = mg + M* + (r^ - r ' J x (V(1) + R) .

Thus, taking the centre of inertia of body Si as the pole we


find, by means of the second equation in (13) and eq. (15),
that
= -mg + ©f1 • u + u> x Of1 • u

(8.10.16)
+ ( Ci -
r T
'K)
X
|^V + Mi [u> x rg + u>x (u> x rg)] j .
440 8. Other forms of differential equations of motion

by using matrix notation and introducing the inverse matrix


(0C) . 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 and in
the plane of the unit vectors , i perpendicular to
2

a = 0,313 + n-iij + o i = ~b a** 2 2

Considering plane , i as the complex plane we introduce the


2

complex number
a* = ai + ia (i = V—l) 2

corresponding to vector a*. Then the complex number ia*


corresponds to the vector

i'3 x a = i'3 x a* = i^ai — i'^,

lying in the same plane. Now, returning to the equations of


motion (13) we rewrite them in the form
M fvc* -\-i3Vc3 + x vc* + 13^3 x vc* — VC3I3 x ^*^
= i'3V3 + V*, ►
+ CM LO + (u>* + i3^3) x (Au>* + Ci^uz) = i' m^ + mf. ,
3 3 3

(8.10.1
7)
Here A and C denote the equatorial and meridional moments of
inertia of the body, so that

©c — A (iiii + 1^2) + ^3^3-


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
M (vc3 -b W1VC2 ~ w2vci) = V3, Ctu3 = m3 (8.10.18)
and equations for the transverse motion
M [be* + i (^3^c* “ ^C3^*)] = V*, (8.10.19)
Aj* -i(C - A) UJ3CJ* = .
These are the equations of motion of the free body of
revolution.
Proceeding to calculating the reaction force R and the
reaction moment
M. we adopt that the surface a separating body S± is the plane
K
8.10 Equations of motion of a free rigid body 441

FIGURE 8.3.
Fig. 8.3. The body Si is also a body of revolution about
symmetry axis i3 with the centre of inertia on this axis at a
distance Isl from C©fso
1
—that
M (ii ii + 1^2) + ^1^3

and
r
cx lq S
■3d?
rfK — i sh.
It follows from eqs. (18) and (19) that
= \ [m* + (c - Azw3w*] .
o;3 (8.10.20)

Substituting this into eq. (15) we arrive at the expression for


the axial force in cross-section a

R3 = -V(l)3 + ^V - MlS (uj + u\)


3
(8.10.21)

and the complex-valued transverse force in this


vw. + ~vt Mls (—imf + CUJSUJ*) .
cross-section
R* (8.10.22)
MA
Before we proceed to eq. (16)
we find
©f1 * Cj — Aid?* C1I3UJ3
= ^ [mf + (C -A) W3I3 x «*] + -
= (Ax - Ci)u3i'zx u>*,
UP X ©f 1 • C b

(rCi _ K)
Y X
MI — -b d; x r'Cl + u x (a; x r'Cl)

1 :/
= Mi (s - h) ( —i3 x V + SUP* + si's x UP*UJ^
442 8. Other forms of differential equations of motion

By virtue of eq. (16) we obtain the torque in cross-section a

TO
M
3= - (1)3

(8.10.23)

and
A\ +the
Mi complex-valued
(s — h) s c M\ bending moment
+------------2-----------+ j f ( s ~ h)zV* +

— \C [Ai -f- M\ (s — h) 5] — CiA\ LU3LU*. (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 mc = 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)
_ r Mi _
V 1 V 0,
- < >+M
since the weight V^) of the separated part Si is proportional to
its mass. Then
m
(l) _ (rCi - 'K)
T xV
(1) - -Jf (rCi - 'K)
T
>
xv

since force V(i) is applied at the centre of inertia C\ of the


body Si.
Equations (15) and (16) take the form

R = Mi [w xr^+ w x (w x rCi)] >


= 0f11 • (1) + u> x 0f11 • d>+ (r'Ci - r' ) x R. K
(8.10.25)

Under a pure translation R = 0 and = 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 = -MlS(a;2+a;2), Mf = 0, 1
p MiC

R* = —^
The transverse force and the transverse moment become
zero provided that the body does not rotate about its
longitudinal axis (^3 = 0).
T n b
ii a cos x + (3 sin \ cosx sinx
*2 a sin x + (3 cos x -sinx cosx

*3 1 a -0

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
a/ = n(/3 + xa)+/3a + rx,
(8.11.1)
the values of the second order of smallness being neglected.
n r ^z> + A sin fjiJ + n ^A cos fi cos v + (i sin vj +
bA cos/x sin z/ + ficosv'j . (8.11.2)
A comparison with formula (2.18.10) yields the relationships
vc •
— = Qb = ~ A cos (i sin u + jx
cos z/, (8.11.3)
JTL
1 = f2n = A cos /x cos v + fi sin
z/, >
where R denotes the radius of curvature of the trajectory of
the centre of inertia of the shell.
The angular velocity vector u? of the shell equals the
geometric sum of vectors u?' and fJ, that is
uj
= r ^ + A sin /JJ + n (j3 + xaj + b - A cos /x sin v + ji cos z^ .
(8.11.4)
Using the tables of the direction cosines we obtain the
expression for the axial component
u>3 = X + v + Asin/x — (3 (/xcosz/ — A cos /JL sin
(8.11.5)
and the complex-valued quantity u>* corresponding to the
transverse component CJ* of vector
u* = e~lxk — \i ^/x cos v — A cos /x sin -f

(3 + ia + i (/3 + ia) ^Asin/x + z>^j e~%x.


(8.11.6)
; . \ .vc
i cos v — A cos fi sm uJ = i—.
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 a 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'l {-a cos x + 0 sin x) + i' 2 (a sin x + Pcos %) + 13]
and, within the adopted accuracy, we obtain
V3C = vc

(8.11.7)
and the complex-valued quantity vc* determining the
transverse component vc* of the velocity vector
vc* = vie + iv2C = vc (—a + if3) e~lx. (8.11.8)

It is easy to prove that projections of vector vc* (but not of vc)


on directions n and b are equal to the real and imaginary parts
of the complex number vc*elx so that
vc* = 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, b and do
not coincide. Also the transverse force F* and moment mf are
assumed to be values of the first order of smallness. Using
raf = pa3e lx [-f[auzvc (~a + ip) - f2iv2c (-a + i(3) -
g[avcA + g^iacu^A].

(8.11.11)
The vector of the gravity force is given by the expression
—Mg\2 = — Mg (rsin/i + ncos/icosz/ — b cos/i sin v).

M ^ruc + n~ftJ =
*3-^3 + F* - Mgi = r (F3 - Mg sin/z) +
2

(na - b(3) F 3 - Mg cos p (ncos v - bsini/) + F*,


(8.11.13)
where the real and imaginary parts of F* represent the
•p
a2 (8.11.14)
M h
446 8. Other forms of differential equations of motion

9 l f f , Jfl
p cos v — A cos p ----cos p cos v ~\~ \ f\ - h ) -TTVCa ~
sin v vc 4
pa" pa uj3 f3 — a ^A sin p M + +
pa |^A cos p sin v — p cos v — a — /3 ^A sin p + , (8.11.15)
92 r
~M
0 = — cos p sin v - (/i - /3) ^rrPvc - fi 2 3
+
Vc M M

paAuj 3 f3 ^A sin p +
91 a + ji cos v — A cos p sin z/ + [3 (A sin +p + v
Mvc
par (3 — a ^A sin p + (8.11.1
92
M 6)
In addition to these, we need the equation of rotation about
the longitudinal axis
. pa4
^3 = — . 9 3 ^ 3 ^ 3 (8.11.17)
and the complex-valued equation (10.19) for the transverse rotation
A (Ae~ixY - i ( C - A ) uj^Ke~ix = m?.
Taking into account (11) we can reduce the latter equality to
the form • C
A - i—cj
A
3A + i (cj3 - x) A =
3
P0b
—j- [/iaw3t;c (~a + ifi) + f2iv% (-a + ip) + SjorocA - g2iauj3A] .
(8.11.1
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
, /?, p, A, v, 0J3, vc-
OL
(8.11.19)
Under such general assumptions on the aerodynamic
forces, the trajectory of the centre of inertia is not a plain
curve since the requirement of vanishing curvature 1/T 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
F * = F * e i x = p a 2( f R + fL)(a-ip)v2c, 1 mf = pa3e~ix
[fMVc (P + «*) - fHave A] , J
8.11 Equations of motion of a spinning447
shell

where
F3 = -pa2fRv2c, m<i = 0. (8.11.21)

It follows from eqs. (17) and (10.18) that the angular velocity
of rotation of the shell about its symmetry axis retains its
constant value c^. Repeating the above calculation we arrive
at the system of equations of motion of the centre of inertia
. P^/R 2 • '
vc =----~M~vc~g sm/i,
. <. g r pa 2

a cos v — A sin v cos 11 = cos g cos v + JL^TI-VCOL, (8.11.22)


vc M
0 = — cosgsin v - /L^-TVCP vc M

as well as to the equation of rotation


A - ^w3A + i (u>3 - X) A = ^-fMv2c (/? + ia) - P°'J‘-vc A. (8.11.23)
The trajectory of the centre of inertia is a plain curve in the
vertical plane A = AQ under the following condition

A = A Q , v — 0. (8.11.24)
In this case, the above kinematic relationship is identically
satisfied and formulae (5) and (6) simplify under the
assumption that g = v c / R , yielding
LO°3 = X , A = i(jt + a - 0 y (8.11.25)
If the coefficient of the aerodynamic lift /L is not zero then,
from the third equation (23), we have ( 3 = 0. This is not
compatible with the equation of rotation (23) since this
equation would take the form
.. , .. . c pa3fM 2 pa4fH
p + a - \ P + a ) = —^—v c a -----------^ — v c ( P + a ) ,
where g is determined by the previous equalities and a is
complex-valued.
pa2 JR ^ -cos/x
vc = — - v c - g s m p , pr -=
(8.11.26)
M vc
define the motion of the centre of inertia independent of the
shell rotation. Then we have the
A
equation for the rotation
a - i p - ^*w3pa
(a -fniff) + p a
V c (a - iff) ~ ^j-fMV2c (a - iP)

Ma2 \ 2g . CUJ°3 :
— g cos g 7rl'*+ —7- f H + - T s m . p -
a (8.11.27)
Avc
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 f 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 motion of the carrying
body. The position of the system can be defined by n
independent generalised coordinates qs, the case of a solid
included. Particularly, 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 unconditionally
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.

A. I. Lurie, Analytical Mechanics © Springer-


Verlag Berlin Heidelberg 2002
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 O and the angular velocity vector
a;. 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 r* and r', respectively. Time is assumed
not to appear explicitly in the expression for r', i.e.
r'i = r - ( g i , - - - , q n ) -
r, =r0 + r ' ( g i , . . . , q n ) .
(9.1.2)
If the motion of the carrying body is given, then ro is a
prescribed function of time and the system constraints are
non-stationary since time t appears in the expression for r*. If
the motion of the carrying body is to be determined, too, then
ro is determined in terms of the generalised coordinates
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
a =a +a; x a, (9.1.3)
$
where a denotes the vector whose projections on these axes
are equal to the derivatives of projections a on these axes. In
dr' .
=r' + w x r ' = 'T-^ qs+u> x r ; (9.1.4)

The kinetic energy of the system is given in Sec. 4.9 by the


sum of three components
T = Te + Tm + Tr. (9.1.5)

Here the kinetic energy of the translation motion Te is as


follows
(9.1.6)
9.1 Differential equations of motion of a carrying body 451

where M and @° denote the mass of the system and the


tensor of inertia at point O, 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 @° and r'c would be constant.
However, in our case, they are functions of the generalised
coordinates q\, . . . , qn. The value of Te does not depend on the
generalised velocities <ji, ...,gn.
The next component is
Tm = v0-Qr + u)-Kr°>
(9.1.7)
where Qr and K£? denote the principal vectors of the relative
momenta and the moment of the relative momenta,
respectively. depends on the values of Vo and u 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 Vo and u on axes
Oxyz fixed in the carrying body as voi and UJI (i = 1 , 2 , 3 , . . . ) . It is
important to notice 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
dr'
Sri = Sr0 + Sr[ Srn (9.1.8)
s =1
Here 0 stands for the vector of the infinitesimal small rotation
of the carrying body and the equality for Sr[ 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
6'W = V • <5r0 + 6 ■ m° + ^ QsSqs,

(9.1.9)
where V and m° 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 quasivelocities voi,Wi. We have
dT^^m^dTrn dT 0Te dTm=

dvoi dv0i dv0i ’ duii du>i duJi ’


since Tr is independent of the quasi-velocities. By virtue of
eqs. (6) and
(4.7.8) we have
dTe
dvQi = M (u0i + u) zc - oj yc), =
2 3

dTe
dv02 M (v 2
0 + 0J3xc - a>izc), = M (9.1.10)
dTe
(v 3 + wiyc - xc),
dVo3 0
and accounting for eq. (4.7.4)
dT
-Q-j- = M (ycv03 — zcv02) + + ©12^2 + ©13^3,
dT (9.1.11)
= M (zCV01 - ycv03) + ©21^1 + ©22^2 + ©23^3,
dT
= M (xCV02 - ycv01) + ©31^1 + ©32^2 + ©33^3-
Here xc,yc, zc denote the coordinates of the centre of inertia of
the whole system referring to axes Oxyz, whilst ©)(. 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
dTm
Qr 1 = MxC,
dTm
Qr2 = Myc,
dTr = Q = Mzc (9.1.12)
dv02
r3

dv0i r '03
and dv
furthermore
dTm _Kr1jy-O dTm dT
~d^~ ’ du>2
K° m K r 3* (9.1.13)
- du3
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 dT dT dT „
"77 Q---1" ^27]--------W3------ — Vi, (9.1.14)
dt dvoi dv03 dv02
d dT dT dT dT dT n
77-7- -h 0J2-Z---^3 ~- - + V02------V03-Z----- = ml > (9.1.15)
dt UUJ1 UUJ2 (JUJ3 UV03 CJV02
9.1 Differential equations of motion of a carrying body 453

since only unchanged three-index symbols for quasi-


velocities Voi,Ui 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
M {l>01 + (ti>2Zc — ^3Vc) + 0^2Zc — ^3Vc) + (^2^03 ~ ^3^02) +
\( J 0 2 (wiyc — U2 Xc) (w 3 Xc — Ml Zc)\ + (x C + U 2 Zc - U 3 yc)} = V L .
(9.1.1
6)
The other two equations are obtained by a cyclic permutation
of the indices. The vectorial form of these three equations
v0 x VQ + tb x r'c + u: x (cv x r'c) + 2CJX rc + r
M c =V
(9.1.1
7)
expresses the theorem on motion of the centre of inertia or,
what is equivalent, the theorem on change in momentum. In
accordance with Sec. 2.11, the value in brackets presents the
absolute acceleration wc of the centre of inertia of the system,
with
we =v0 x VQ + cb x r' + ljx (cj x r'c)
c

(9.1.18)
being the translational acceleration composed of the
acceleration of the pole
and centripetal and rotational accelerations
wc = CJX (to x r'c), wr = dJ x r'c,
(9.1.20)
respectively.
Since the axes are rigidly bound to the carrying body their
angular velocity coincides with that of the carrying body and
thus
Lb = +0J X UJ =LO . (9.1.21)
The term in eq. (17)
wgor = 2u;xrc

Wj- = Vc (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)
*' dr'r dr'r <is + d2r'
dqs dqkdqs^s (9.1.24)
r
°=£^' ^2
s=l k=1

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 (ycv03 ~ zcv02) + ©n^i + ©12^2 + ©13^3 + cu 1©^ + ^©12 +


^3©13 + + M [iU2 (V02%C ~ ^01 Vc) ~ ^3 (^C^Ol ~ %CV03)] +
^2 (©31 + ©32^2 + ©33^3) — Ll>3 (©^^1 + ©^CJ2 + ©23^3) +
~ WZK?2 + M [^02 (wiVC ~ ^2Xc) ~ Vos (^3 C ~ Zc)\ = \ •
X rn
(9.1.2
5)

The corresponding vectorial form is as follows


*0 *o
©° • & +0 -u> + u x 0° • u;+ Kr +
UJ x + Mr'c x ^v0 +u; x v0^ = m°.

(9.1.26)
In order to transform this equation we use definition (4.3.20) of
the tensor of inertia ©° to obtain
0 o
r\° (T? / dri 1 / dri 1 ®ri /
=21?.^^ (9.1.27)
s=l i= 1 dqs 2 ldqs 2 dqs

Accounting for eq. (4.9.6) we can write


O nN
'r / dr'
0 -co + co x K(? = 2^g^ i•
S=1 2=1 . dr*
!dr' , 1
„/ dr( K
r, • w--r l • cjj+-ux r, x dq
2% 2 0qs
s
or, after simplification,
2X>f>r'x(cox|l
0 -to + to x KY = 5—
N
2 1 = 1 ' *
G( S
/
2j2mi*i x (ux .
(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
Cor — ~ 0
m
•<*> - (JJ X K^\ (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
0 • (JJ + (jj x 0 • u = mc + m£or — Kr .
C C

(9.1.30)
Here 0 denotes the moment of inertia of the system at its
C

centre of inertia
0C = ©° - M (Er'c • r'c - r'cr'c) .

(9.1.31)
The resultant moment of relative momenta about the centre of
inertia is
= K(? - Mr'cx rc

(9.1.32)
and the moment of the Coriolis forces of inertia about the
centre of inertia is
m
*c
Cor = ~ 0 -w - u; x K(9.1.33)
Finally, mc denotes the resultant moment of the active forces
about the centre of inertia
mc = m° - r'c x V.
(9.1.34)
Of course, eq. (30) can be obtained directly from eq. (26) if
pole O 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) .
456 9. Dynamics of relative motion

9.2 Differential equations of the relative motion of carried


bodies
Required are Lagrange’s equations for coordinates qi,...,qn
determining the position of the carried bodies mounted on the
carrying body. Due to eqs. (1.5) and (1.9), these equations
are
&s (Te) + Zs (Tm) + £s (Tr) — Qs (s = 1,... , n).
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
dT dr' ld@°
£s (Te) e = —M (vo x LJ) • dq 2 dqs
(9.2.2) s
dqs
We proceed now to the second component. Using eqs. (1.7)
and (1.8) and taking into account that vector VQ does not
depend on q and q we obtain
s s
drc
dQr
£ (vo • Qr) = VO •
S + V0 • £s (Qr) = -Mv0dq• + Mv o • £s s
dq s

By virtue of eqs. (1.3.5), (1.3.11) and (1.3) we


have
d tc _ (ir). d ihJc _ dr'c _ drc dr'c (9.2.3)
dqs dqs ’ dt dqs dqs dqs dqs
and thus
d drc drc dr'c
dt dqs dqs dqs ‘
Hence,
£s (vo • Qr) = M (v0 +U> X V0) • + M(v0xw)' dr'c , * dr'c
dqs dqs
(9.2.4)
Vector uj does not depend on q and q , too, therefore we obtain by analogy
s s

£, (w • K° u:-^+u-£s (K?) v r/
dqs \dt dqs dqs J
) dqs
dK?\* dK° dK.?'
. dKp ' = u> • dqs ) dqs dqs
- +u ■
dqs

But
dK°
u • u; x dqs =0
9.2 Differential equations of the relative motion of carried bodies 457

and the obtained expression can be written in the form


£,{u-K?)=u.^+u,-£;(K°),
(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 a
cast
f in the form
£s (Tr) = Qs ~ M (v0 + 0J X V0 j +
dq
1 <90° s
— UJ • - (9.2.6)
2 Oq

We turn our attention to the terms on the right hand side of


this equation. The vector
S° = -Mvo = —M (v0 +u> x v0^

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
0 s° • = -M (vo +v X Vo) dr'
Qs dqs

^M(^0+tJXV0)-r/c (9.2.8)

is termed as the s — th generalised force of inertia for pure


translation, whereas the scalar product
n° = M ^v0 +u x v0^ • r'c
(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
= (9.2.10)
can be referred to as the potential energy of the centrifugal
dUu _ 1 <90°
dqs ‘1° dqs (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 expression
N f n N r, f
Kr° = 5>ir'x ri= J2is'Emiv'i x Ji
2=1 s=l i= 1
we obtain
dK°
"• ■at=“'!><r:xfji = f>‘("xr:)'£-
2=1 2=1
The vector
SI = -mi (ti x r') (9.2.13)
represents the rotational force of the inertia of the particle.
Thus, the quantity
2=l2
(9.2.14)
dqs dqs
is the s — th generalised rotational force of
inertia. Returning to expression (12) we
have
zdK^y V N
^qk^rrii ( (dr' dv' k , 82r'
dqs ) S= 1 \dqk dqs + Ti x
x

dK° 2
= qk mi ‘ X + r?: X (dqs " dqk
d r'i
dqs dqkd
S=1 2=1 qs
so that

c* (TTO\ O dv'i dr'i


■ £, (Kr ) = -2u, ■ x —. (9.2.15)

This expression can be represented in either of two forms: the


first form
N irixdr'4
(K?) = Q? = -2u>.'£m
or
2=1 m - v. V
dqs
(9.2.16)
2=1
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

• £*s (K?) = Ts = J2 IskQk


(9.2.17)
k= 1
with the gyroscopic coefficients (7.3.6)

1
-* = 2“'|>‘gjxai = '
(9 218)

Using the introduced notation, the system of differential


equations for the carried bodies is written in the form
r\
£s(Tr) = Qs- — (n° + n") + <# + r8 (s = i,.(9.2.19)
The left hand sides of these equations depend only on the
quantities describing 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 qi,..., qn and six first order differential equations for
the quasi-velocities vok and u>k- 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

9.3 Relative equilibrium


When considering the relative equilibrium of carried bodies
determined by the equalities
qs = q°s = const (s = 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 u). 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 relative velocities qs and accelerations qs equal
zero. Thus both the kinetic energy of the relative motion Tr and
the gyroscopic forces Ts become zero. In addition to this,
according to the adopted assumptions on the motion of the
carrying body, the generalised rotational forces of inertia are
equal to zero also.

where, if Qs depends on the generalised velocities qs, the latter


should be replaced by zeroes. In eq. (2)
n° = Mw0 • r'c, IF = -|fa> • 0° • w. (9.3.3)

Vector r'c and the tensor of inertia 0° depend on the required


coordinates
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 </i,..., q-m and
the cyclic coordinates gm+i,..., qn. Then equation

£m+l (Tr) — rm+/ + Qm+i (/ — 1,... , n m), (9.3.5)


9.3 Relative equilibrium 461

where Qs, II, II0,IP', 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
Es (Tr) + Es (Tm) + Eg (Te) = Qs — —— (s = 1,... , n).
OQs
Provided that WQ and UJ are constant, the motion under which both po-
sitional coordinates
Qs=q°s (s = l,...,m) (9.3.6)

and the cyclic generalised velocities


qm+i = <im+i (1 = 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 nn
£s(T ) = J2 ttskqk T EE [k,j;s]qkqj k=1
r k=1j=1

and the definition of Christoffel’s symbols, that for the


positional coordinates

n—m n—m
e. (Tr) EE [TO + l, m + r; s] qm+lqm+r (9.3.8)
1=1 r= 1
n—m
iv—iivn—miv—iiv r-v
I ~ ^ ^ Qm+lQm+r (5 — 1, . . . , 171)
dqs
v

,1=1 r=1
whereas for the cyclic
ones
['m + l,m + r;m + t] qm+iqm+r = 0
EEn—m
n—m
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 n—m ^ m m+r o
o X] Yl ------ »l’ Qm+iqm+r + Qs ~ (II + II0 + IF) +
r\ 'imtnmi-r ' ^ s r\
1=1 r=1 oqs
dqs
n—m 1=1
0 7m+l,m+2 •
' ' • 7m+l,n
7m-
Dn—m — f2,m+l 0 ..' • 7m+2,n (9.3.12)
<
7n,m+ 7 n,m+2 • ■ • 'Inn
< /
1 ■

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, determinant (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 Dn-m = 0.

9.4 Equilibrium of rotating flexible shaft


We consider an elastic rod with round cross-section whose
end O rotates with constant angular velocity uo 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.
ii h 13
ii 1-j(/J + 7 )
2 2
j+-a/3
2
-(3 + -ay
i'2 l-i(7 + «2)
-7 + a+
1 -j(«2+/j2) |
*3 /?+ -a +
^f3y
464 9. Dynamics of relative motion

Replacing vector UJ by a;i3, due to eq. (2.10), we obtain

IR ari3 • 0° • i3.

We have
i3 • (ErM • rM - rMrM) • h = u 2 + v 2,

13 ErM • r'c
- - (rMr’c + r'crM) ■ 13 = rM ■ rc - i3 • rcrM • 13
= (uii + vi2) ■ rc + H3 ■ r'c - i3 • r'cl = (wii + vi2) ■ r'c.

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
(Mil + V\2) ■ r'c = ux'c + vy'c + (x'cv - y'cu) + z'c (u(3 - va).
Let @ik denote the components of tensor 0 M referring to the
axes Mx'y'z' fixed in the body, then with the help of eq. (4.4.11)
we obtain
i3 • ‘ 13 — 011^i3 + ©22^23 + 033O33 +
20i2^i3a23 + 2023a33O23 + 20i3ai3a33
or, by means of the table of cosines,
i3 • 0 M • i 3 (3 (022 33 o?
(0ii — 033) 2
+ — © ) + 033 —

2012OLp + 2©23 (ol + —^7^ + 2©I3 + -0^7 ) .

Now by virtue of eq. (2.10) the potential energy of centrifugal


forces is as follows
n" = —-a;2 [(On — 033)/?2 + (022 — ©33) ~ 2012^/? + (023/?+
©13a) 7 + M (u
2
+ v 2) + 2Mz'c (u(3 — va) + 2M7 (x'cv - yrcu)\ -W2
[023a - 013/3 + M (x’cu - y'cv)\ - ^033^2,

(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 Il e of the elastic strain
nci = = \i Lydz =
1
| {- vf
2EIJ 2 El u
0
l K
- / L%dz
1
ne2 = = 1 i f- °:
2 El J 2 E' y
0

>i!i
'2 y1y

/2
I-Vy°L°
2
V X

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
„ 1 12 El u2 + v2 - 2^ (u/3 - va) h (a2 + 2) + i^72. (9.4.5)
"'=2 —

The latter term corresponds to the potential energy of


torsion. Neglecting the potential energy of weight we arrive at
the equations of equilibrium
(ne + nw) = o,
dqs
466 9. Dynamics of relative motion

which, in the expanded form, take the form


u / 12El \ _ „ f 6EI £c \ y
l \M13LJ2 ) P
\Ml3w2 l) l1 l'
12EI QEI
v
( d | a( \ 'c \ 'c
z x
V'c
l \ M13LJ2 \ Ml3w2 l l l’
AEI 9n — Q33 \ Q12 _
0 M13UJ2 Ml2 ) + MPa~
U / 6El Zq \ 1 023 _ 013
> (9.4.6)
AEI 022 — 033 \ 012 o , M/2 Ml '
l \Ml 3 2+
u l ) 2 7- 2

Ml3u2 ' Ml2 ) MPP+


V / 6EI 1 013 _ 023
l\Ml3u2 + l ) 2 Ml ~ Ml '
2l 2

uyb _ 023 g 013 C _

llll 2MEP 2MI2“ M I V 7 - ’

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 Mzf being its axis of symmetry. Then
©11 = ©22, ©12 = ©23 = ©31 = 0, xc — yc — 0,
and the latter equation says that 7 = 0, i.e. the rod is not
twisted. We introduce the notation
3 El ©11
_ — ©33
Ml3u2 Ml2 ~ - (u + iv) = p, a + i(3 = e.
(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 parameter 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 < ryj3 and k > 0 for h > r\J3 where r and h denote the radius
and the height of the cylinder, respectively.
We obtain the system of homogeneous equations
(Au - 1) p + ie (2u + Cc) = °>
2 2
1
-i (2u + Cc)P+ (jX
2 2
-^e = 0 j ^9'4’8)

with the determinant


A (u2) = (Au2 - 1) ^v2 - kj - (2v2 + Cc)2 = | (^4 - av<1 + b)>
(9.4.9)
9.4 Equilibrium of rotating flexible shaft 467

where
a—3 ^~4^_ • (9.4.10)

If the determinant (9) is not zero, then, in the equilibrium


position e = p = 0 the axis of the rod is a straight line. The
angular velocities which are roots of equation A (z/ 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 + IT' reduces to
the single inequality A (z/2) > 0. 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.
a2 -4b Cc + ^ + & (*-&) (9.4.11)

4
( Cc + 2) + 6 (Cc + 2) (^ 3 ) + 3 ( *
-

It presents the quadratic form in (c + 1 and k — -j- and since 42 • 3


— 3 >0
^ o ^
this quadratic form is positive definite becoming zero only
when (c = — -
and k — —. The roots of equation A (z/ 2) = 0 are always real-
valued. Only one root is positive if
b < 0 or Qc>k.

(9.4.12)
When
b 0, CL > 0 or k C,Q-) C,Q + - + & > ( ),

(9.4.13)
o
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)
Cc + Cc + g < 0,
2
(v + 4) (9.4.15)

2
(v - 4) (9.4.16)

under condition (13).


The case of a single critical velocity uo\ occurs for any disc
since b < 0 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 = 0 and ()C = 0, i.e. a particle on the shaft
is considered then a = 1, b — 0 and as relationships (7) and (9)
show, the critical angular velocity UJO is equal to the frequency
of free oscillation of this mass. For a thin disc we can take C c
= 0, then
z/2 = I (\ + 3k+ ^(1 + 3k)2-3k^j ,

and as k < 0 it is easy to prove that the

inequality

5<"5<1
holds (it is necessary to2take
> into^account
= that -4
\/9 k2 + ...
> « — 3k
I-
(9.4.17)
wl v{
The gyroscopic effect increases the critical angular velocity 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 UJ\ and UJ\ corresponding to
the roots v\ and v2 occur under condition (13), that is for ”
cylinders”. If denote the smaller root of (9) then, as eq. (16)
shows, the straight form of equilibrium is stable if J1 < UJ\ or J1 >
u)\ and is unstable if J <
1
< UJ\.

9.5 A gyroscope in Cardan’s suspension mounted on a


moving platform
We introduce three trihedrons of unit vectors: es,is,ig bound to
9.5 A gyroscope in Cardan’s suspension mounted on a moving platform 469

FIGURE 9.3.

angular velocity <£, is mounted on the inner gimbal. The


bearings of the axis of rotation of the outer ring lie on axis e 3
of the platform and those of the gimbal on axis 12 of the outer
ring, so that e3 = i3, i2 = if2. The axis of the rotor rotation
coincides with the unit vector i' l5 see Fig. 9.3. The origins of
all axes coincide and are placed at point O 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
rf = aiii + CL2I2 + ^313- (9.5.1)
The acceleration of point O which is taken to be constant
both in value and direction is denoted by WQ and the constant
vector of the angular velocity of the platform by <*?. In order
to take into account the force of gravity, we enter the vector
of the geometric sum
W* = w0 - g,
(9.5.2)
with g being the acceleration due to the gravity force. Then,
we set in eq.
(3.10) n* = np + n°
n* = raw* • r',
(9.5.3)
IIg and n° 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
r = r0 + r', v = v0 + u? x r' + r*'.
ei e2 e3 ei e2 e3
ii cos a since 0 ii cos (3 cos (3 sin — sin/?
h — since cosce cos ce ce
0 ii — since cosce 0
is is sin (3 cos sin (3 sin cos (3
0 0 1 ce ce

as well as the following expressions for the vectors of the


relative angular velocities of the outer ring, the rotor and the
gimbal
w'j = e3d, '
u)’n = —i^d sin j3 + i3d cos f3 + i2(3 = e3d + i2(3, > (9.5.6)
u'm = u'n + i[<p, ,

respectively.
We proceed now to construct equations (3.10) and begin
by calculating the derivatives of II* with respect to a and /?.
By virtue of eqs. (39, (5) and (1) we obtain

an* duj'r
da = mw* • da x r | = m (w* x e3) • r',
II

an* Mi
= mw* • (\2 x r') = -mw* • (i'3ai - i[as).
9(3 = mw* • dp x r
9.5 A gyroscope in Cardan’s suspension mounted on a moving platform 471

Denoting the projections of w* on the platform axes by wswe


have
w* = w iei + w2e 2 + w3e3,
then, by means of the tables of the direction cosines, we
obtain the relationships
<9
11* —mw\ (ai cos P sin a + a2 cos a + a3 sin P sin a) +
da mw2 (ai cos P cos a — a2 sin a + a3 sin /3 cos a),
(9.5.7)
<9 —ma\ (wi sin (3 cos a + w2 sin (3 sin a + rc3 cos (3) +
11*
ma3 (w\ cos (3 cos a + w2 cos (3 sin a — sin (3). j
We turn our attention to construction of the potential energy of
the centrifugal forces. The tensors of inertia at point O of the
outer ring, the gimbal with the additional mass and the rotor
are given respectively by
Op — C/iiii + Bi\2\2 + A/i3i3,
Op/ = [Bn + m (a§ + af)] i2i2 + [An -(- m (af 3- a2)] *3*3+
[C/j + m (a| + a§)] iii'i - m [aia2 (i'ii2 + i2i'i) + (9.5.8)


«2«3 (i'2i3 + 1312) + s i (i3ii + iii3)],
a a
Here A and B denote the moments of inertia about the central
axes lying in the mid-plate of each part and C about the axes
perpendicular to this plane. As a result, the expression for
takes the form
r (0/ +0/7+0///) • Bi (w • Ci (a; • ii) +

i2)2 + Ai (U> ■ i3)2 - 1 r ./ \ 2


^i) +
B (u> • i2) + + A (u; • i3) — 2maia2u •
2 2
• i^> —
2raa2a3u; • • i3 — 2ma^a\UJ • i3u; • i^] ,
where the constants A, H, (9.5.9)
C can be easily expressed in terms
of the values entered by formulae (8). The differentiation of
scalar products w -i's is performed by means of rule (5), i.e.
du-¥s dcv'j
11 x 1 = (w x e3) • i's = (cj ei -
da = uo c^ie2) • i's, =
da
2

du> ■ i' du'j11


dp = LJ x 1 ^ • (i2 x i') (s = 1, 2 , 3 ).
dp
(9.5.10)
472 9. Dynamics of relative motion

As follows from eq. (3.10) we are interested only in the


coefficient of p2 in the expression for the kinetic energy Tr of the
relative motion. This coefficient is constant and equal to \Cm
and thus the corresponding component in eq. (3.10) drops out.
In order to calculate the gyroscopic coefficients 7a(p and 7^, we
notice that in this particular case eq. (3.11) is an identity, i.e.
— 0. According to eq. (2.18) the following expressions
N
f)r’ f)r' N
dr < dr\
= "
/ cap
2
E mi
~dp x dip'
i=1 2=1
are required to be found where the summation should be
performed over all particles of the system. Turning to eq. (5)
we have
da dp d^'in dp
da x r, x
du'm A du'jjj du'm x di^'ui \
0a 7 dp da dp )

and moreover
N du'in du'iu
7aV = 2w]Tmir'r'
2=1 da x dp
N (9.5.11)
7/3*, = U-J2 i i i
2 mrr
d(3 du'm
2=1 X dp

Now, recalling definition (4.3.20) of the tensor of inertia we find

mTr
i i i = E mir'i ' ri ~ G° = (0n + &22 + ©33) - ©°
2=1 2=1 (9.5.12)

and particularly, for the rotor which is a symmetric body,


N
1
^ ra*r'r' = -E (2Am + Cm) — [A///E+ (Cm ~ Am) i^i'J
2=1
= -j^Cm + (Am - Cm) i^i^. (9.5.13)

Thus we have
loop = \pinw + 2 (Am - Cm) LO • i^ii] • (e3 x i[) = Cmu • i^cos/?, 1 l(3ip =
\pinw + 2 (Am - Cm) w * ij.ii] * O2 *1) = —Cmu> • J
x

(9.5.1
4)
9.5 A gyroscope in Cardan’s suspension mounted on a moving platform 473

and furthermore
Ta = CUICp (—sin a + uj2 cos a) cos /3,
(9.5.15)
T p = —CmCp (uq sin 0 cos a + UJ2 sin 0 sin a + UJ3 cos 0).

These results can be anticipated since expressions (15)


obtained are nothing else than the projections of the vector of
the gyroscopic moment
T = CmCpii x a;, Ta = T • e3, Tp = T • \2 (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 0 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 e i , e 2, e 3 are taken as being
directed along the axes of the geocentric system: ei pointing
north along the tangent to the meridian, e 2 pointing west along
the tangent to the parallel circle, and e 3 along the upward
local vertical. Then, denoting the northern and eastern
components of the vector of the platform velocity as VN and VQ,
(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 centrifugal 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 U2 are

(9.5.18)

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
a\ =0, (12 = 0, as = —a
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, . \ •a • ^2 \ o'
— (wi sm a — ic2 cos a) sin p = — sin a —— cos aJ cos p,

— (w\ cos a + W2' sin


((jJla) cos (3U)2
—— . \ . n , ^3 n]
sin (3 q cos a + — sm aj sm p + — cos pj ,
(9.5.19)
where
CinCpU
9 = (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
q—a---------(3 qu 2
U9 u’
(9.5.21)
W 2 . w3 u;i \ =u;i o;3
- - -a + — + 9jT)P --------------977-
9 ^ U Q U
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 VNof the rotating
VQ earth (9.5.22)
RU cos RU cos 4>
5 5

respectively, are small compared to unity.


The ratios of the components of the Coriolis
acceleration to g
2VNU cos<3> 2V<JU 4>
(9.5.23)
COS

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 RU2 /g 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 atVNthe formulae /3 = —q sin4>, (9.5.24)
a=
RU5cos
4*
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
VNof order (22) then VN Vp ), (9.5.25)
a — RU cos VO RU cos RU cos
1+'
4> RU cos 4> 4> $
9.6 Relative motion of rigid bodies 475

1 +------ - - - q sin4> ^ ——------ — q cos .


/? = —q sin4> RU cos 1 +i?[/ cos 4> /
1 + q cos (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 O
are bound to the carrying body specified by index 1. The
origin of the axes C2x'y'z', fixed to the carried body marked by
Mr'c — Mir'Ci + M2r^2,
(9.6.1)
where y denotes the position vector of the centre of inertia of
!
c
the system introduced by eq. (1.17), whilst r'Ci and r'C2
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 qi, q2, #3 describing the position of the centre
of inertia with respect to axes Oxyz and three angular
coordinates #4,(75, qe describing the orientation of the
ifc = ife (?4,95,9e) (k = 1, 2 , 3 ), (9.6.3)
with i'k being the unit vectors of the axes of the carried body.
Since the position vector r'Ci is a constant vector relative to
axes Oxyz, we have the following obvious equalities
M rc= M2 r( c2> Mr n= M2 r c2> (9.6.4)
where, due to eq. (1.24),

c2 =
r dr Ci ** /
£ dqs Qs, c2 = E
r (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
Mi jv0 +u x vo + cl? x r'Ci + LOX (OJ X r'Ci)
vQ +CJ x v0 + d? x r'C2 + u> x (<JJ x r'C2) + 2ux rc<2 + r c
= V.
(9.6.6
)
Along with vq and a;, this equation contains the generalised
coordinates QUQ2, Q3 and the corresponding generalised
velocities and generalised accelerations.
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 equality
©° = ef + ©o

where ©j* denotes the tensor which is constant with respect to


axes Oxyz so that
*o *o*o
0i = O, © =©2 (9.6.8)

Furthermore, it is natural to introduce into consideration tensor


0^2 denoted for brevity as ©2- By virtue of eq. (4.4.2), we have

©? = ©2 + M2 (Er^2 • r^2 - r^J


(9.6.9)
and using the formula for differentiation of tensor (4.3.13) we
obtain

Here <*?' 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 O is equal to

= r'C2 x M2 rC2 +©2 • u/,

(9.6.11)
so that its time derivative with respect to axes Oxyz is given by

Kr = y'c2 x M2 *r Ca +02* w +o/ x02-w/.


9.6 Relative motion of rigid bodies 477

for eq. (1.26)


^ ^ L/
0 -<jj + <jj x + Kr = a/ x 02 • a; — 02 • (a;' x UJ) +
*/ r */
Cc? X 02 ’ Cc? + 02" 0^ ~\~UJ X 02 • Cc? -|- A/2 2r^»2*
I*£<2 Cc? —
. * f fr
* ( , \ , **7 (9.6.13)
r
c2 rc2 •"- rc2 rC2'W + wx (rC2 x rC2 J+ rC2 x r
C2

By means of eqs. (7) and (9) we have


0° -W + W X 0° • U> = 0^* -W + W X 0® • U> + 02 • X 02 • U>+
m
2 (r'c2 ■ ' ^ - w x rc,2r'C2 • a;
)
and
ci/= c5 +CJ x a/.
(9.6.14)
Now substituting the latter equations into (1.26) and using
simple transformations we arrive at the following vectorial
©f • UJ + UJ X ©f • CO + Mir^ X W0 + 02 • (d? + u/) +
(uj + a;') x 02 • (uj + a;') + M2r'C2 x wc2 = mo- (9.6.15)
Here, for the sake of brevity, the following notation for the
vectors of acceleration of the pole O and the centre of inertia

w0 =V0 +Ct?o X v0,


Wc2 = W0 + UJ x r'C2 + wx (UJ x r'C2) + 2u;x Tc<2 + r* Ca (9.6.16)

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 0 • b — b x 0 • a = (0 — #E) • (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
'd = ©ii + 022 + 033?

(9.6.18)
’ d *< dv'C2 *>
Jt - '
rc
dqs °
2
dqs )

^c2 , d dr'C2 *> d*c2


+
dqs (wx rk) • dr
'c2,*' Sq, dt dqs °
2
dqs

Making use of transformation (2.3) we find as expected that


Ss (Tr) = M2 r'Ca dr'C2 (9.6.24)
dqs ‘
9.6 Relative motion of rigid bodies 479

The equations of motion (22) take the form


dr'
M2WC2 --^■ = Qa (s= 1, 2 , 3 ),

(9.6.25)

where the vector of absolute acceleration wc2 °f 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 94,95,96- The calculation order is
dU° dU" dq8 ’ 1 d©
—(jj -------2 LC (s 4 , 5, 6 ) . (9.6.26)
dqs 2 dqs
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 u' is a linear combination of the generalised
velocities q^, q$, % 6
o/ 6

u duj . _ (9.6.27)
E • 9s s—4
/
s=4
. 9qs ^
with the coefficients
duj
es
' (9.6.28)
dqs
depending on the generalised coordinates # 4, 95, 96- For
example, if Euler’s angles are taken as the generalised
coordinates, i.e.
94 = ^ 95 = $, 96 = if,

then vector u;' is represented in the form


u/ = In$ + i'3(p,
(9.6.29)
where i3, n, i!* are the unit vectors of axis Oz, the nodal line
and axis C^z', respectively. In this case

du' duf duf


e
-777- — 13, e5 = —7- = n, e6 = —- = i3.
4
<?94 dq$ dq6 *
Applying formula (27) we can cast eq. (10) in the
form
* V-^ <702 ^ _ .
©2=s=4
l_j -frj-is s—4 ( s X ©2 - ©2 X
e

e 3)
480 9. Dynamics of relative motion

and thus
——- = es x 02 — @2 x es.
(9.6.30)
dqs
The generalised centrifugal force has the form

Q“ = • (es x @2 - ©2 x es) • = (a; x es ) • 02 •

(9.6.31)

Furthermore using expression (11) for vector K£? we obtain

by eq. (2.14)
= -w • =02 • e,.
(9.6.32)
UQS
Determination of the generalised gyroscopic forces reduces to
calculating the Eulerian operator over vector K^?, i.e. turning
to eq. (11) we have
C(K?) = (e2-es)*-^-02-u/
= ©2- es + 02 es - • u/ - 02
(9.6.33)
dqs dqs
*
Thus we need es . Applying formulae for differentiating the unit
vector of the coordinate axes of the carried body
dqs ^ k> \dqs
since vectors i'k depend only on the generalised coordinates.

The result is

^ x i ^ + w ' x (es x i'fc) =es xi'fc + es x ( u / x i'fe).


Accounting for the identity
es x (a/ x i'fc) - a/ x (es x i'fc) = \'k x (u/ x es),
we can recast the above result in
the form
do w' x es x ij. = 0, k = 1,2,3,
/
9.6 Relative motion of rigid bodies 481

yielding
* duj
r
,
es= —----|-w xea, s = 4 , 5, 6 .
----------(9.6.35)
Using formulae (10), (30) and (17) we can transform
expression (33) as follows
£*s (K?) = 02 • (u>' x es) + («' x 02 - 02 x w') ■ es -
(es x 02 - 02 x es) • ui' = (202 - EI?2) • (es x u / ). (9.6.36)
In addition the generalised gyroscopic forces are

rs = 2w • (©2 - 1E^2) • (<*/ x es) (s = 4,5,6).

(9.6.37)

The gyroscopic coefficients are, due to eqs. (2.17) and (27),

7sfc = 2w • (©2 - |E02) • (efc x es) (k, s = 4 , 5, 6 ) .

(9.6.38)
We arrive at the following differential equations of rotation of
the carried body
£s (Tr) — Qs + (^ x es) • ©2 • LO — (b • ©2 • es +
dw' i , oe2
£s (Tr) = (es ■ 02 • a/)* - — • 02 • «' - -a/ • —±
dqs 2 dqs
*'
= es • ©2- a; +es • (a/ x 02 - 02 x a/) • u/ + (a/ x es) • 02 • a/ - •
(es x 02 - ©2 x es) • a;' = es • ^02- u? +u/ x 02 • .
(9.6.40)
• LJ ~\~LO X ©2 • CO -f- U? X ©2 • tjJ 4“ ©2 * UJ4-

2u/ x ©2 (9.6.41)

Replacing here d> by CJ according to formula (14) and using


identity (19) we arrive at another form of equations for the
rotation of the carried body
es • [©2 • (tb + ib') + (u> + <J) x 02 ■ (w +1*/)] = Qs (s = 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
02 • (ct? T u;) + (CJ T ci/) x 02 • (CJ T CJ ) — m^2,
(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 62-
Varying the generalised coordinates 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
Referring to eq. (5.2.5) we can conclude that
6
y QsSqs = m°2 ■ O', Qs = m°2 ■ es (s = 4, 5 , 6 ). (9.6.45)
6’=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
*f
derivation is simpler, for example, there is no need to
calculate u: , only a/.
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 rc<2 and a/ on axes C2x'yfzf
*' +'
c2=rc2 +o/ c2> LsO =LJ LO
xr +'
r +u/ X
1
(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^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
Mo vr +a/ x vrT v0 +u; x v0 + Co x r'C2 +
u: x (u; x r^2) + 2u x vr] = V2, (9.6.47)

*f
where vr =rC2 is the vector of the relative velocity of point C2.
In the equations of rotation we should replace Qs and es by
projections of the resultant moment m°2 on axes of the carried
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
02 • ~b CcJ ^ -f- UJ X 02 • OJ ~b U)' X 02 • Cc/ ~b
2OJ' x ^02 - • u> = m°2.
(9.6.48)

9.7 Examples

9.7.1 Equations of rotation of the 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 ei,e2,e3 of the geocentric
coordinate system described in Sec. 9.5. The origin O, 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 O, that is
0° - Ai[i[ + BI'2I'2 + Ci'3i'3.

(9.7.1)
The angular velocity vector UJ of the carrying body has both
constant value and direction. One should set w in eq. (6.48).
The projections of w on the axes of the carrying body and the
axes of the geocentric system are denoted by u)8 and cjs,
respectively. Thus
3
uj\ = U cos<4>, c<;2 = 0? ^3 = U sin<f>,

(9.7.3)
$ 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
J x [(A - B - C) i'iii + (B — C — A) i'2i'2 + (C - A - B) i'3i'3] • w =
(C - A - B ) - L O ' 3C J 2 (B - C - A ) } i' + ( A - B - C ) -
u [ u 3 ( C - A - B ) } i '2 + KC52 (B-C-A) - u'2 o>i (A-B- C)} i'3.
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
AuJ^T ((7 — L?) (uj2uj3 T ^2^3) T A (002^3 — (D3CJ2) H~
(C - B) ((JJ'2CJS + UJ3QJ2) = mi, (9.7.5)
and the others being obtained by a cyclic permutation of the
letters and indices.
The terms of the type 0)20)3 originating from the centripetal
forces of inertia are proportional to U2 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
Auj'y T {C — 7?) (uj2uj3 T uj2uj3 T0^30)2) TA (0)20^3 —
0)30^2) — mi,
Buj2 T (v4 — C) (0^30^ T uj3uji T0^0)3) TB (0)30^^—
0)10^3) — ui2, >
Clo1 T (B — v4) (o^^o^2T ujf^uj2 H~0^0)1) TC {uj\lo2 —
9.7.2 Heavy top
As an example of the application of the 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 Oz' by angles a and /? described in example 7.18.2
and apply the system of axes n, n',^ ’’halfbounded” to the
heavy top. As usual, (p denotes the angle of rotation about axis
Oz'.
Equations (6) are written under the assumption that axes
are fixed in the body and, for this reason, the second
Ox’y’z'
equality in eq. (6.46) was used for the derivation. Now this
equality should be replaced by

where LO denotes the vector whose projections on axes n, n', i3


are equal to the time derivatives of the projections of vector a/
on these axes (denoted by Lj's as above). The vector of the
angular velocity of the ’’half-bounded” axes A differs from a/ in
the absence of the spin component, i.e.
A = J - i'3<p.
Now
+ ' -(pm! + (p\\<jj'
LO + ' 13(f LC -L
- X
O 2

and the quantities in eq. (6) should be replaced by + (puo2,


Co + CpCj^Co3.
2
9.7 Examples 485

The role of axes £,77,^ 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
00n = Co 1 = U (— sin <£> sin f3 -f cos 4> cos (3) = U cos ($
+ /?), con> = Co2 = —U sin (<£> + /?) sin a, = 003 = U sin (<£>
-f /?) cos a.
is
m° = — i3z' x esQ = Qz' (—nsin/3 — n'sin a cos (3 + ig cos ce cos/?) x i'3
= Qzf (n' sin (3 — n sin a cos (3).

Accounting that A = B, differential equations (6) become

In the case of a nearly vertical top, the angles a and [3 as


well as the projections uo'^uo^ of the angular velocity vector are
small. Also 00s being proportional to U are small. Neglecting
the products of small quantities, we reduce the first two
equations (9) to the form
A (ciq + Cpoo^ T ^2^3) T (C — A) (002 + Co2) ^3 — —
Qzfa, A (Co'2 — (pco[ — LOIOO3) — (C — A) (oo[ +
Coi) oo'3 — Qz'{3,
and, by virtue of the third equation,
oo'3 — const.
Within the adopted accuracy of calculation we are to accept that
Co 1 = U cos $, Co2 = 0, co3 = U sin <£>,
= -a, co'2 = f3, uo'3 = Cp.
The system of differential equations (10) can be now written
in the form of a single linear second order differential equation
3
(P+ ) - p (f + iaj — Qz' ((3 + ia) = C(pU cos <1>.
i& Ci(
A
(9.7.11)

It has a particular solution

0 = 0o = (9.7.12)
486 9. Dynamics of relative motion

corresponding to deviation of the axis of the top to the south


(if Cp > 0 and z' >0). This motion is imposed by oscillations with
frequencies
Ai \_C<p
(9.7.13)
4QAz'
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
Cip > 2^QAz'. (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 O. Let the
axes Oxyz be bound to the carrying body. The flywheel rotates
with angular velocity u/ = (fi's 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 T'c.
In eq. (6.6) we have now
*'
v0 = 0,
r
c2= 0j wc2=wc=ux(wxr'c) + iixrc.
The term x wc is transformed as follows
OJ x M2 (Er^ • r'c - r'cr'c) ■ u> + M2 (Er'c • r'c - r'cr'c) ■ u>.
Let 0° denote the tensor of inertia of the system consisting of
the carrying body and the ” frozen” flywheel at point O, then
0° = 0? + M2 (Er^ • r'c - r'cr'c) + 0f.
Using eq. (6.15) we can recast eq. (6.15) in the form

©^ • UP UJ x ©^ • UJ T ^^13 x ©2^ • I3 T 2^13 x ©2^ • u;+


<P (©!? - E0? ) • (i' xw) + e^ (^3)* = rn°. (9.7.15)
Since the flywheel is a body of revolution we denote its
central moments of inertia by ©1 = ©2, ©3, to get
0f = 01E + ( 0 3 -0i)i&.
9.7 Examples 487

Thus
*3 x ’ *3 —
• (^3)* = 0^ • (<pi3+ X ig) = 03^3 + 01^ X ig

and
2^3 X ®2 ‘ & + if (®2 ~~ ' (*3 x ^) + ’ (^3)

= 2Q\(p\3 x UJ -\- Cp [0^ — (20i T 03)] 13 x UJ T 03^bi3 4- 01 CpLO


x i3
= 03^3 + 03Cpu x i3.

(9.7.16)
Equation (15) is rewritten as follows
0° • Co + u; x 0° • u + 0 3 (^3 + (poj x ig) = m°.
(9.7.17)
0° = -Aiiii + Bi2\2 + C^is.

Then projecting eq. (17) on these axes we arrive at three


differential equations
+ {C — B) (J2UJ3 + 03 [aip + ~ U3P) <p\ =
AUJ1 '
BUJ2 + (A - C) CJ3CJ1 + 03 \(3ip + (uo3a - (jij) <p\ = (9.7.18)
rriy, >
CUJ + (B - A) UJ\UJ2 + 03 [lip + (ui/3 - 012®) <p] = m°, y
where 3a, /?, 7 denote the cosines of the angles between
direction i3 and axes Oxyz.
The equation of rotation of the carried body, i.e. the
flywheel, is constructed in the form of eq. (6.41). Under the
adopted notation
duj'

and furthermore
ig-0^-i = ig • [0iE + (03 - e^igig] -ig^ = 03^),
13 = iji • [0iE + (03 - ©i)^] -u = 03 {wla + w2p + d>3l)

The other termsin the brackets in eq. (6.41) form thevector
which is
orthogonal to i3 and do not appear in the equation for rotation
03 (ip + <jia + L02/3 + ^37) = Qp, (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
AuJi + (C — B) LU2(^3 + ^ ©3 [Pkak + Pk (—27k ~ —3Pk)] = mx >
k=1
n
Bd)2 + (A-C) U3W 1 + X ©3 [iPkPk + ‘fk (u3ak - wi7fe)] = m° >
k=1
n

CUJ3 + (B — A) U)1U)2 + X 03 [v>fc7fc + <Pk (^1 Pk - u2ak)\ = m°


l- -1 >
(9.7.20)
and correspondingly
©3 (V>k +uiak +d}2(3k + W37fe) = Q<pk (fc= 1,... ,n).

(9.7.21)
An expression for the kinetic energy of the carrying body
with the flywheel obtained in Subsection 4.12.2 is given by
11 1 T = -u; • ■ u> -h - (u> -h if3(pj • ©^ • (u; + i3Cp) + —
M2 |u; x r^l
= ■ [©? + ©2' + M2 (Er'c • r'c - r^r'c)] • u> + ^©3<£2 + w • 0^ •
= • 0° • (Jj + i©3^2 + (c^lCK + V20 + -37)
<£•
In the case of n flywheels

T = -u; • ©° • u + - ^ ©3 [(p| + 2 (aqa*; + a;20k + -37,) <£,] •


(9.7.22)
It is natural to take the moments of the active forces about
axes Oxyz in eq. (20) as independent of the rotation angles ipk of
the flywheels.
The generalised forces Q^k 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 dT applied to the flywheels. The coordinates cpk are
Pk = -Q^r~ = ©3 [<Pk + 2 (-!<*, + W2&k + 1,... ,n).
-37,)] (9.7.23)
The Routhian function, due to eq. (7.9.13), will have
the form
i?=Me°.uMx0; Pk l 2 (9.7.24)
lq| - (wiafc + U}2f3k + u3Ik)
k=i
9.7 Examples 489

and yields the Euler-Lagrange equations

ACo 1 + (C — B) LU2U3 + ^2 {(^27k ~ ^3Pk) \_Pk — ©3 (^l^fe + w2/?fc


+
k=1
^37fc)]+afc [<2Vfc =mI° (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=QVk (k = l,... ,n).

(9.7.26)
If ipk 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 determining in terms of the
d_
n
1 Alo\ + Bid2 + Cu)3 — ©3 (id 1 QLk + ^2Pk + +
dt 2 ^37k)
n
53 Q<Pk (W1 k + v20k + 0J3jk) =m°LUi+myU)2+m°uj3
a

k—1
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 n of eqs. (26) and (23), as follows
7-1 n 2- n
Q<pk 0^1 ak+U2pk+^37^) .41 y ' 2k. Qvk&
k=1 dt 2 0S k
k= 1 n^ k=1
dR k—1
o
dt
Turning now to formula (7.9.15) we can obtain the power of
the active forces
n 7
A'' = y^QVk<Pk + ^2 =
^ (^2 - #0) •
(9.7.27)
fc=i
The example of a body carrying flywheels demonstrates
clearly the concept 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 (carrying body) as the pole O, we can write the
differential equation of rotation
+
’W + WX ^©C + ^0C'i^ • tjj +
N
y |^©Gi • u) +u;' x Q • a;' + 2a;' x @ • u-
i
Ci Ci

(6 ©n + ®22 + ) w i x w N
+ y^rriiY^ x wi = mG. (9.7.28)
i= 1

Here Y\ — GGi stands for the position vector of the centre of


inertia C* of the i — th carried body. The position vector Y' of the G

joint centre of inertia of the shell and the carried bodies is


determined by the equality
N N
M + ^2mi ) r'G = ^2miTi, (9.7.29)
i=1 i=1

and the absolute acceleration of point G is equal


to
w G = w c + d; x r ^ + o ;x ( u; x r ^ ) +
N
1 **f\ — (2u;xri+ r (9.7.30)
i=1
i=1

Let us express mG on the right hand side of eq. (28) in terms


of the resultant moment mG about the centre of inertia G
N
mG = m G + Y' x V = mG +
G rrikj Y'G xW G

N
mG + x w
[ c+ Co x YG + o;x (a; x r^)] +
fe=1

rb X £>, (2a,x + r^) .


(9.7.31)
i=1
9.7 Examples 491

Inserting the latter equation into eq. (28) and replacing w* by


its expression we obtain
N N N
© + ^2© ’w+wx (© + X ® • + X ® ■
c Ci c Ci w x Ci w
2=1 2=1 2=1
@Cz • uji +u;' x QCl • a;'
— + @22 + ©33)
+
N
X X (r' - rjj) + a; x [w x ( r ' - r'G)]} +
2=1
(r' - r'G) x (2a; x r* + r •) = mG (9.7.32)
2=1

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\ of this body is given by
r' = TQ + aee (t), (9.7.33)
where the mean value of function e (£), within a sufficiently
long time interval, is zero. Then rg determines the mid-
position C® 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 mG = 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 M that m r', u/ = 0, (9.7.34)
r'
r — r - rG =
G— M+m M+m

and the equation of motion (32) can be written in the form

(0c + OCl) -U; + U) x (0c + 0Cl) -w


Afm
(9.7.35)
+ m r' x [u x r' + u x (a; x r') + 2aeoj x e + aes].
M
We replace here r' by its expression (33), put
the term
Mm , r , , ,.,
M + mr0 x [(jj x r0 + (jjx (a; x r0)J
onto the right hand side and introduce the tensor
©Go + qGo = QC + QC! + Mm (Er[) • rg — igi-g) , (9.7.36)
M+m
492 9. Dynamics of relative motion

which is equal to the inertia tensor of the shell and the body ”
frozen” in the shell at the joint centre of inertia Go- Indeed, by
eq. (4.4.2)
0G° = ec + M( Vr'Go-r'Go-r'Gor>Go),
qGo = ©Cl +m [E (rQ—r£j0) • (r'o-r'cj - (rj,—r^0) (r^-r^)] ,
and we arrive at eq. (36) by adding these equations and using
formulae (34).
Equation (35) can be rewritten as follows

(©G° +ef°)-w + wx (©G° + 0G°) • w = LG°

(9.7.37)
Vector ~Lg° can be referred to as the perturbing torque about
LGo = 2 Mm
as (t) Ero • e - 2 (roe + ero) u>+
M+m
Mm
oox 0’ - 2( o o)
Er e r e + er
• 00 aV (t) [(E - ee) • <b-
M+m
2 Mm
w x ee ■ w - as (t) (Er^ • e — erg) •
M+m
2 Mm 2
a e (t) e (t) (E — oo
ee) • oo—
Mm ae (t) rg x e. (9.7.38)
M+m M+m
In this expression only the terms with non-zero mean value
are essential. They are

= — —------a2eLV2 f(E — ee) •y oo — oo x ee • oo]J.


~L*GO (9.7.39)
M + ra
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 oo and oo 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
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 4>i (£,
77, £) which is a harmonic function in the coordinate basis 0£r](.
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 transformation of the coordinate systems we can express
the velocity potential 4>i in terms of x, y, z. Clearly, time will
appear
Thus, explicitly in this dependence
va = grad 4>,

(9.8.1)
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 = (v0 + u) x r') • n = v0 • n + UJ • (r' x n),

grad (v • r') = v ,
0 0 (9.8.3)
as vector VQ does not depend on the coordinates x,y,z.
We enter the harmonic vector B (x, y, z) that is the vector
whose projections Bi, B2, L?3 on axes Oxyz are harmonic
functions. They are determined within the cavity V by means
of their normal derivatives on surface S
dB1 dB2 dBs
-^-=yn3-zn2, = zm - xn3, =xn2-ynx. /n0/A
(9.8.4)

Here ni, n2, ns denote the projections of n on axes Oxyz and


dBs dBs dBs dBs
dn dx U\ + dy n2 + dz ■n3 (s 1,2,3)
denote the normal derivatives of functions Bs.
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
/ dBs
(9.8.5)
dn do = 0
494 9. Dynamics of relative motion

holds. Here do denotes the surface element and integration is


over the surface S 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

S V
(9.8.6)
where dr 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 S. In our case
dy dz
_ dy dr = 0,
s dz
which completes the proof.
Conditions (4) can also be cast in the form
n • gradB = r' x n.

Here we introduce tensor gradB which is dyadic product of


the operator
. .d .d .d
grad = ii— + i2 + 13 ~TT~
ox oy oz

and vector B. Notice that in particular gradr' = E, E being the


unit tensor.
Given vector B, the velocity potential 4> can be found by the
relationship
4> = v0 • r' + B • (jj. (9.8.8)
It is a harmonic function of x, ?/, z as r' and B are harmonic
vectors whereas Vo and u are independent of the coordinates.
Time appears in 4> 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 4> satisfies the boundary condition (2). Indeed,

and then, using eqs. (3) and (7), we find


n • va = n • grad 4> = n • v 0 + (r; x n) •
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
/ *f
a — v0+u;xr + r ,
v

and comparing with eq. (9) yields


r = grad (B • a;) — UJ 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
M [v0 + (jj x v0 + us x (cj x r'c) + u x r^] = V, (9.8.12)
with M denoting the mass of the system.
*o
The tensor of inertia 0 of the system at point O remains
constant in
eq. (1.26) for the same reason. It coincides with that of the
system with
the ’’frozen” fluid. Thus
*o
0 =0
and all we seek is the moment of the relative momenta and its
time-
derivative. By virtue of eq.
v (11) we have v
(9.8.13)
Here pdr = dm denotes an element of the fluid mass and p its
mass density. Furthermore
) J r' x (a; x r') dr = uj • J (Er' • r' — rV) dm = uj • 0 (9.8.14)
v v
496 9. Dynamics of relative motion

is the angular momentum of the ’’frozen” fluid about pole O


and ©^ denotes the inertia tensor at point O.
Now we proceed to calculation of the integral
G° = p J (r'x gradB • u) dr = + ©£ • u>.

(9.8.15)
v
As follows from eq. (4.8.8), vector G° is the angular
momentum of the fluid in the vessel provided that point O is
fixed.
Denoting the tensor of inertia of the rigid body (without fluid)
at point O by
and noticing that
*o o
Kr x = —@2 * d; — UJ x ©^ • u;+ G +UJ x G°,
we can write eq. (1.26) in the form
©? • W + u> x 0f • u> + MY' x (v0 +u> x v0) + G +w x G° = m°.
C

(9.8.1
6)
Let us find the projection of vector G° on axis x. In principle,
we could operate with vectors but this would complicate the
derivation. We have
<»*»>

V
The integrals are easily transformed by means of formulae (6)
and (4) to the following form

J (y~Q^~ ~ dT = JBk (yn3 - zn2)do = JBk~§^do' (9-8-18)


V v
S' s
Then introducing the notation
dB'
/
~g^Bkdo (i,k = 1,2,3),
(9.8.19)
5
Gx — QllVl + Ql2UJ2 + Q13U3 (9.8.20)
and two analogous expressions for G® and
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 harmonic functions we obtain
p dBi R , _
Q?k f Bkdo P dBj 3Bk dBjdBk dBi dBk
dr
s— on + dz dz
dx dx dy dy
'J dBk
s dn Bido. (9.8.21)

This also means that Qfk = Qki.


Projections of vector G° are linear functions of projections
of vector CJ. Thus quantities Qfk 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 O since the angular
momentum of the fluid about point O can be represented in
the form
G° = Q° • u,

(9.8.22)
similar to the angular momentum of the rigid body having the
tensor of inertia Q° at point O.
Equation of motion (16) is now cast as follows
(0? + Q°) • Co + uj x (0f + Q°) • u + Mr'c x ^v0 +u; x v0^ = m°,
(9.8.2
3)
whilst in the case in which the fluid was ” frozen” it would have
the form
(0f + 0^) • Co + u x (0f + 0^) • u: + Mrfc x ^v0 To; x v ^ = m°.
0

It is known that
- -— -11 dF 2z
711
A dx a A’ U2 A dy b2A’ 713 A dz c2A’
2

where
dF\2 dF 2
A= _
\dx J dy J dz
498 9. Dynamics of relative motion

The first of these boundary conditions is written as follows

and is satisfied by the following harmonic function

By analogy, we obtain

2 ,
.....
Calculation by eqs. (21) yields

B, Qg =

Here A, B,C denote the moments of inertia of the fluid in the vessel
A=\M (b +c2), B=\M (c2 + a2), C = \M (a2 + b2).
2
000
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
[98]. spherical vessel.

9.9 Equations of motion for a solid


In this section the carried bodies are the particles of an elastic
solid vibrating 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 VQ of
the pole and the angular velocity vector <*? are bound to this
carrying body. The position of point M of the skeleton is given
by its radius vector OAt = p and TQ = OO denotes the radius
vector of the pole O relative to the fixed axes with the origin at
point O. 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
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 freedom 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,qi,... ,qn) (9.9.1)
is given. This expression can be represented as a power
series in terms of
qoc
n in n

u=X (x’ y>z) + 2 X X z


)+■■ • (9.9.2)
06=1 06 = 1 (3=1
Here we took into account that u =0 at qa = 0.
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
u0= X hqa<Pa (X) + X 13qn+a^a (x) ,

(9.9.3)
06=1 06 = 1
where the vibration modes in planes xy and zx can be taken as
ipa (x) and ipa (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 72 and 73, then
n n
1272 = -12 Qn+atPa (x) , I3I3 =06=1
06=1 13 (X ) ‘
The additional component of vector u, which determines the
displacement of the point of cross-section with abscissa x and
Hi = (1272 + 1373) x (i2V + hz)

-ll y X] (X) + ZY1 (X) (9.9.4)


. 06=1 06=1 .
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

£ = x- dx,

where £, 77, £ denote the projections of the displacement of


the rod axis on axes Ox, Oy, Oz, respectively. It is also
assumed that £ = 0 at x = 0. Making use of eq. (3) we obtain
n2
^'oc 0*0 + Qn+ai’a (x) dx.
jDt = 1 ot=1
Therefore, accounting for the terms quadratic in the
generalised coordinates we obtain the following expression
for vector u
U = ^2 { [ w'c, ( ) + i2Va (a?)] + <ln+a [~iiZip'a (x) + i3ipa (a;)] } -
qa -il x

a=1
nn x x
5‘> EE
a=ip=l qaq/3 J <p'a (Ov'p (0 + qn+aqn+p J i>'a (0 i’p
(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 practical 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 integral characteristics which are not
sensitive to these functions provided that their choice is
reasonable.
In what followsr'=
thep integration includes
+ u = xii + yh not only distributed
+ zi + u. (9.9.6)
3
9.9 Equations of motion for a solid 501
The position vector of the centre of inertia C is given by the
equality
Mr'c = Mp c + Judm = Mp c + &(q 1 ,... ,qn), (9.9.7)

so that
(9A8) d2a

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 carrying body and eq. (2.19) for coordinates
qa. These equations must be linear is qa and qa. 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 potential energy II of the inertial forces in a pure
0

translational motion, and the potential energy IT*' of the


centrifugal forces. Moreover, it is necessary to take into
account the terms linear in qa in the expression for variation 6rf
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
f *' *' f* *7 . . f du du
2 Tr = r • r dm = uu dm = } } qaqp ------dq- - -dm,
J J J « dq«
and it is clear that for calculation of T it suffices to keep only
linear terms in eq. (2). Then we obtain
nn «
Aa0 Aa0
^
Tr = OL=l = U“ • U^' (9.9.9)
(3=1 ''

Turning now to eqs. (2.5) and ( 8) and cancelling out the terms
which are independent of qa we have
II = M ^v x v0^ • r'c = ^v x v0^ • a
0
0 0

(v0 +W v0)
. (9.9.10)
= X

a=l a=l (3=1


Here the notation
d 2a \ Va0dm (9.9.11)
dqadqfji = i
502 9. Dynamics of relative motion

is introduced. Thus
an0 /3=1
Q0
a dqa ( V0 +U> v0 )
X
(9.9.12)

and neglecting the quadratic terms in formula (2) would lead to


a loss of the linear terms in the expression for QQa.
The tensor of inertia 0° which determines nw by means of
eq. (2.6) should be found with the same accuracy. Due to eq.
(4.3.20), we have
0° = J (Er' • r' — r'r') dm

=J [E (p + u) • (p + u) - (p + u) (p + u)] dm

= J (Ep. p- pp)dm + 2 J Ep • u-i (pu +up) dm +

/■ (Eu • u — uu) dm,

and, in order to keep all terms which are quadratic in qa, we


should take into account such terms in the expression for u.
Then we obtain
0° = 0? + 2^Aa9a + ^^QaV9/3- (9-
9.13)
a=l a=l0=1
Here 0|p is the tensor of inertia of the rigid skeleton at point O and

Aa / Ep • UQ — - (pUQ+Uap) dm,

Qa/3 J EUQ • - 1 ^“U^+U^U0) +

Ep • UQ/3 - 1 (pUQ/3+UQ/3p)| 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
F = F* + J2 (fQ<?« + (9.9.15)

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 hydrostatic 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.
Sr = Sr0 + Sr' = Sr0 + 6 x r' + S*r (9.9.16)
n \ n/ n '
p+J2<i*va + y^ u^ + ^Tir^
( ot=l / 0=1 V 7=1
v
6'W = J F • Srdr = J, <5r0 + 6 x (p+y^U“<zQ +
Ot = 1
dr +
a=1 7=1
[J 11, (f/3^ + f,/3^) ■ (Sr° + 0 X p + y] \Jaqa j dr.
0=1 V a =i / (9.9.17)

The terms quadratic in qa and qa have been omitted here and dr


denotes an element of the volume, surface or arc.
Let the resultant vector and the resultant moment about pole
O of forces F and F* be denoted by V, m° and V*,mSp,
respectively. Then the elementary work is given by
n
6’W = V-6r0 + m° •0 + '52Qa6qa. (9.9.18)
a=1

Here

V = V* + Il(q0 f f 0
dT + q0 f t^dr) = V* + £ (*/% + V%) ,
0=1 ' JJ p=1
(9.9.19)
nrr «
m=m ? + T/ \(10 J [p X f,J + Ui3 x F*J dr + Qg J p x f,i3<iT
0=1
= m? + y^ (*A/3 + v'0q0) ■ (9.9.20)
0=1
504 9. Dynamics of relative motion

The constant vectors vf(3, fiP, \il^ 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 Qa in eq. (18) take the form
Qa = f F* • Uadr + J2 q0 f (f0 • U“ + F* • UQ/3) dr+
J 3=1 L J

(3=
qp J ?? • Uadr = Qoa + X] + r'a0Qd) ■ (9.9.21)
/3=1
The term Qoa 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 qa are zero in the natural
state, the expression for the potential energy of elastic forces
is a quadratic form of these variables
nn
ca(3qaqp-

(9.9.22)
Z
a=l(3=l
Taking into account formulae (7) and (8) and notation (11), we
arrive at p + p )
M following equation of motion of the carrying body (1.17)
the
V0 +U> X V0 + X c X (w X c

n
= V* + ^ ("“4“ + V>aia)
a=1

— ^ {qa [u> x a“ + u> x (u> x a")] + 2qau> x aa + qaa“} . (9.9.23)


a=1
The case qa — 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
where
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

0° •w + wxK?] =-2 f^qa (u ■ A“ + ha x GQ


' OL-1 '
It is easy to prove that the above equality is the moment of
the Coriolis forces. Indeed,
wAa + lwxG“ = -|px(wxU“) dm,
where this equation is obtained by means of eqs. (14) and
(25) for Aa and
G°7
Substituting the obtained equalities into eq. (1.26)

we obtain 0^ • CJ + CJ x 0^ • CJ + Mpc x ^v0 +cc x v0^


n n
= m? + (naqa + n'aqa) - ^ qa [2 (AQ • w + w x AQ • w) +
a= 1 o;=l
n n
a x (v0
a
x v0^ j - ^2 Qa (2a; • Aa + a; x Ga) - ^
(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 of rotation and
the generalised gyroscopic forces Ta by means of eq. (2.14),
"7r'xi:‘,m=“7(p+g®u'’)
<£ = -

U“ + y]u7“g7 \dm = -<b • GQ + Gfaqp (9.9.28)


K 7=1 J \ (3=1
where
g/5« = j ^ xU“+px
dm. (9.9.29)
Expressions for the gyroscopic coefficients 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 qp. This yields
/ — x —dm = 2OJ • V
a
T a (3 'I' (3a
= 2lo J 9qa &Q(3 J x U0dm = 2u> •
rQ/3
(9.9.3
Z/j J .z

1
I /y
TIT —i—y~*
\ ____L-J_ _ °z y
, . ,. ... .j
1

FIGURE 9.4.
and equations (2.15) in the form

('r<+ r/a%) - Y ca% -


Y Aa(3qp = Y
(3=1 (3=1 (3=1
fa n \ In
a + ^2aQ/3<?/3 j ’ (vo x v0) + u> • I Aa + ^2 Q0,0qp w+
0=1 (3=1
n
2w-£r [ G “ + £G%j] (a 1, ,n).
(3=1
3=1
(9.9.3

1)
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 aa, Ga, Aa 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

9.10 Oscillations of a rotating rod


This problem serves as an example of the application of the
general equations (9.31). Bending vibration of the rotating
elastic inextensible rod (a blade) is considered. The end O of
the rod is clamped in a turbine wheel which rotates with a
constant angular velocity UJ about axis Ozi, 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
9.10 Oscillations of a rotating rod 507

the axis of the undeformed rod. Then


i x,
p= 1 CJ = cji3, v0=cjRi2, CJ x v 0 = -oj2Ri1 (9.10.1) and neglecting the
weight we can cast eq. (9.31) in the form
n n
E ^ + E ( ° ~ “2Ra? - OJ2Q^) q0 +
Aa0 c 0
(3=1 (3=1
n
2a^rf*(j/3=u,2(A“ +a«R).

(9.10.2)
(3=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)
UQ = —iiy<p'a (x) + i2<pa (X), Un+Q = -iizip'a (x) + i3tpa (x), '
X X
= -ii J q>'a {o ^ (o de, un+“-"+^ = —ii J ip’a (o ipr0 (o d$, >
0 0
(a, (3=1 J
(9.10.3)
where
U“,,3+n = \j
a+n
'd = 0 (a,0 = l,... ,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
i
J f (x, y, z) dm = J p (x) dx J f (x, y, z) do.
0 S(x)
As functions <pa (x), (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 ipa (x) and (x), then under the above
conditions
•Pa (*) = •Pa (*) • (9.10.5)
Hence, for the case in question
U“ = hfa (*) , U"+“ = Wa (x) , '
X
(9.10.6)
Va0 = Un+a,n+l3 = _h J ^ (£) ^ (£) ^

which reduces the volume of calculation by half. Using eq.


(9.9), we obtain
l
Aoc(5 n+*,n+{3 = pSiA^^ *(3 = / 'Pa( )'Pf3( ) - (9.10.7)
s s ds
=A A

0
Here and throughout this section the integration is carried out
over a nondimensional variable s = x/l.
As follows from eqs. (9.11) and (9.14) the projections aj of
a** on axis Ox as well as the components A33 of tensor 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
1 s
= a^+ ’ +0 — —pS J ds J ipfa (a) ipfp (a) da.
a n
(9.10.8)
0 0
Turning to eq. (9.14) we obtain

1 s
Q\ da ds,
j <pa (s) <p0 (s) - s J ip'a (a)
= pSl
0L 0 ► (9.10.9)
1 s
V; da.
^33 = ~PSI J sds J ip'a (a)
tp'p (a) 0 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 U a and Un+Q: have different directions
so that the values EUfc • Us - 1 (UfcUs + UsUfc)
13 13
9.10 Oscillations of a rotating rod 509

in coefficients Q33 (k,s — 1,n) differ from those for fc, s = n + 1,2n.
The potential energy of the bent rod is calculated by means
of the well- known formula
l L
wff dx,
Ue = ±EIZ J v"2 (x) dx + 1EIy J

where v (x) and w(x) denote projections of displacement of the


rod axis on axes Oy and Oz, respectively, and Iy and Iz stand for
the moments of inertia of the cross-sections about axes Oy and
Oz, respectively. In our case

n
e = I ^2 + Qn+aqn+l3^?r')J f v'L 00 ‘Pfi («) ds,
a=lp=l ' ' 0
(9.10.10)
which yields the coefficients
i
(9.10.11)
J p'L (») Pp («)

The mixed coefficients with one subscript less than n and the
other greater than n are absent in eqs. (7)-(ll). One can easily
see that the gyroscopic generalised forces vanish in this case.
Indeed, turning to formulae (9.30) and (6) we obtain

-1 j3cx
=i3
7 ud x Uadm = 0. (9.10.12)

We enter the notation


>2 _ EIZ 2 _ ±-,ly EL, (9.10.13)
pSl4 ’ pSl4

1 s
m n+a,n+0 = J (j + s)ds J ^ ^ ^ ^
a/3 = m
da.
0 0
By means of the following simple formula
ay
J dx J f (x,y)dy = j dy j f (x, y) dx,
00 00
(s) - "iVa (s) = 0 (9.10.17)
subject to the boundary conditions

Va(0)=^(0) = 0, V "(1)=<(1) = 0. (9.10.18)


Here
A2 (9.10.19)
4

\a denoting the a — th natural frequency and


- El A PSl*'
(9.10.20)

so that A2 — A2 and A2 = A2 for vibrations in the plane Oxy and


Oxz, respectively. The values of isa are known from the
frequency equation and equal
Vi = 1.19372—, 1/2 = 0.99610—, i/3 = 1.00010^,
Z Z ZJ ► (9.10.21)
2k-1 , „
Vk~ „ TT, 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

(9.10.2
2)

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)
l
A“0 = J (fa (s) tpp (s) ds = 6af3, (9.10.23)
0
c
a/3 J Va ( ) <^3 (») ds = <p'a (s) (a) 11~ J ‘f'a(s)‘P0
s

ifi'p 0 (s)ds 0

[Pa (S) ¥$ (s) - <Pa (s) Vp (s)] |o + J <Pa (s) V™ (s)


0
and, referring to eqs. (17), (18) and (23), we find
Cat3 = 6a0vAa.

(9.10.24)
The equations of dynamics (15) and (16) are considerably
Qa +
(X2a) +UJ2 qa+u2 ^2 ma0q0 = O (a = l,...,n) (9.10.25)

0=1
for the tangential vibration, and the form
n
qn+a + {xi)zqn+a + U!2'^2ma0qn+I3 = 0 (a = l , . . . , n ) (9.10.26)
0=1

for the axial vibration. Here ( A a ) a n d ( A a ) ^ 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 ma(3 determine the restoring effect of the
centrifugal forces. If we omit the quadratic term in the
512 9. Dynamics of relative motion

FIGURE 9.5.
u we would not obtain the terms with coefficients maf3 and the
physical meaning of the problem would be lost to a great
extent.
Normalised expressions for functions <pa (x) satisfying eq.
(23) are given by
Va (x) = -7 [(sinhz/Q, + sin va) (cosh vas—
sin cosh va — cos va sinh va
cos vas) - (cosh + cos va) (sinh vas - sin uas)],
so that, for a given value of i?/Z, the numerical calculation of
coefficients maf3 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 permanent 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 consideration. The carried
bodies are the gaseous combustion products in the
*/
nozzle of the rocket. Vector is the velocity of the i — th particle
relative 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(X).

(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 = l implying
the inlet and outlet cross-sections of the nozzle, 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
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'c — 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
* ft/ j|c/\ * o r **/

0 -u + UJ x = 2 / r x Lx r J dm, Kr = / r'x r dm. (9.11.3) After

these replacements eqs. (1.17) and (1.26) take the form M v0

xv0 + £ i x r c + wx(wx rc) + 5|C / / 5K


/
r dm + r dm = V, (9.11.4)

0° • a; + a; x 0° • a; + Mr^ x (vo + a; x vo) +


2 J r' x ^a;x r ^ dm + J r'x r* dm = m°. (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
N N vi _d_ N 9
OY i d 0
dq E
V
i QS
2—1 dt dq ^m,2 ;
dq mi
s s
T
2=1
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
0°. 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 m* with respect to time t. However one
can forget it and erroneously take time-derivatives of the
integral characteristics like M and 0° relevant to the kinetic
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

where M denotesasthe
follows = —Mdt
flowdmrate and= vthe
(A) mass ,
density
pSdt = pSdX p is
assumed to be constant. For this reason, taking into account
eq. (1), we obtain
i i
Jr dm =- e J v (A) pSd\ = eM Jd\ = Mle. (9.11.8)

By virtue of eq. (1) we have


r -ev' (A) A (9.11.9)
and, furthermore, applying
eq. (7)
J r' dm = Me iJ v' (A) Adt = Me J v' (A) dX = Mv (l) e.

(9.11.10)
00
Here v (l) denotes the value of the relative velocity of the gas in
M v0 +u; xvo+wxr^+wxjwx r'c) = V + $ + FCor. (9.11.11)
Here $ denotes the thrust
§ = —eMv (l).

(9.11.12)
Its value is \M\v (/) = —Mv (/) and its direction coincides with that
of e as M <0. The effect of the motion of the gases in the
nozzle results also in the Coriolis force
FCor = _2u, x Me L (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
r' = r' + e (/ - A),
(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 O of axes Oxyz
r dm = eMdA, r* dm = Mev' (A) d\. (9.11.15)
9.12 Gyroscopic platform 515

Hence
2 J r' x x r ^ dm
i
= 2Mr[ x(wxe)l + 2Me x (a; x e) J (l — A) dA
o
= 2M x (a; x e) l = — J + x FCor (9.11.16)

as well as
/ /
Jr'x*r'dm = M J r[ x ev' (X)d\ + M J (l-A)ex ev' (A)c?A
0 0
= Mv (l) r[ x e = —rj x

(9.11.17)
We arrive at the equation of rotation of the rocket

= m°0°+ •r[cbx +$ cb
+ x( 0°
r{ +• UJ + Mr
] xc FxCor
(v0 + u? x vo)
f
-eZ
(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
$ 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

9.12 Gyroscopic platform


The carrying body is a platform with mounted gyroscopes
which are gimbals with rotors. The axes Oxyz with the unit
vectors ii, i3, i3 are fixed in the platform, the angular velocity
vectors and the velocity of pole O are denoted by u and vo,
respectively. An orthogonal trihedron of unit vectors a*,, b*;,
Cfc = x is bound to each gimbal, vector 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
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
an<
Xk ^ Vki then the angular velocity vectors of these bodies with
respect to axes Oxyz are
u
k = akXki uk = akXk + (9.12.1)
see Fig. 4.3 taking into account the change in notation. The
derivatives of vectors a^, bfc,c*; referring to this coordinate
system are given by
5|C
a
^ ^
fc= 0, bfc= u'k x bfc = XfeCfc, ck= u>'k x cfe = -bfcXfe-
(9.12.2)
Hence
*U / .. .. . U = a
k— k kXk + bk&k + CkXk&k- (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'k denoting the position vector OOk.
The masses of the k — th gimbal and rotor are denoted by m'k
and mk, respectively, and their inertia tensors at point Ok are as
follows
where, for brevity,
Ak = Ak + Ak, Ck = Ck + Ck, rrik = mk + mk. (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 2 s (s
gimbals and s rotors).
Let m and r'Cp = 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 v0 +uj x v 0 + u ; x r c +u;x (w X rCP) +
fc=i +CJ x v0 + Co x r'k + cv x (u x r = V.

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 O is taken
in the centre of inertia of the platform and masses rrik at points
then
mv'Cp + mk*'k = 0 (9.12.6)
fc=i
9.12 Gyroscopic platform 517

and the equations of motion for the centre of inertia is


M (v0 x v 0) = V,
(9.12.7)
where M is the mass of the system. By the appropriate
attachment of balance masses one can make point O
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 at point O is given by

©° = ©p + mk (Elfc ‘ r'k ~ rfcrfe) • (9.12.8)


k=1

Using equalities (3), (6) and notation (5), (8) we arrive at the

equality

j 0 ° + E ^ J 4 f c ) .d, + d,x&°.u, + 22(Ck-Ak)(bku>-bk+


V k=1 / k=1
s
LO x bkui • bfe) = m° - {Akak\k + [Ckoj x ak +
k=1
2 (Ck — Ak) Cfcbfc • ui)xk + Ck (ipk bk + c k&kXk + ^ x bkftk)} •
(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

Tk =^ + S-kXk) ' [EAk + (Ck — Ak) bfcbfc] • (u> + akXk) +


- (a; + &kXk + bk<Pk) ’ [^4k + (Ck — Ak) b^b^] • (cv + akxk +
bk(pk).
Transformation yields
Qxk — Q<pk ~ 9, (k !,...,,§). (9.12.11)
518 9. Dynamics of relative motion

Coordinates ipk are cyclic. However Xk are the positional


coordinates, as vectors are dependent on them. We can
construct s equations which express that the cyclic momenta
(the kinetic moment of the k — th rotor) are constant. The result
is
Cji(<pk+ufbk) = Hk (k = l,...,s) (9.12.12)
By means of eq. (7.17.2) and (7.15.4) we construct the Routhian function
2 [^k (Xk + ^Xk^• ak + w2) + (Ck — A'k) (u: • bfc)2 +
Rk
U
2Hku • bfc - ^ (jfe = l,...,s)
kJ

and the differential equations


£Xk(Rk) = Ak(xk+u-aik)-(Ck-Ak)u>-bkck-LO-H<jj-ck = 0
(k = l,...,s).
While deriving the latter equation the following relationship
dbk _ d bfc _
dx dx °k
was used. Now it is easy to prove eq. (9) which is the
vectorial form of the Euler-Lagrange equation for the
Routhian function
s
1
R= -w0° -u:+Y^Rk.

(9.12.15)
The first of these equations
d dR dR OR _ c dt dui + W2
dwz W3
du>2 ~
mi

takes the form


(@° • u;) + LO2 (@° * ^)3 — <^3 (Q° • ^)2 + [Ak (Xkakl +
^l) +
k=1
(C*. - Ak) (u> • bkbki)* + Hkhl + AkXk {u2ak3 - ^3^2) +
(C*. — Ak) u) • bk (w2bk3 - <^3^2) + Hk (u2bk3 —
^3^2)] — m?• It represents the projection of the vectorial
relationship
0° + ^ j • u; + u; x 00 • LO + ^ ^ {Ak&kXk 4" AkXk^ x afc4~
k k=1 / fe=l
(C* - Afe) [u • bfcbfc + u; • bkv x bfc + (u • cfcbfc + u; • bfccfc)
xfc+
Ffc(xfccfc + WXbfc)]} = m° (9.12.16)
9.12 Gyroscopic platform 519

on axis Ox. Equation (9) can be used to deduce this form if Cpk
and ipk are replaced by the corresponding expressions from
eq. (12) and the following equality
u x ak = v • ckbk -w • bkck = a; x (bk x ck)
is applied.
The rotors are assumed to rotate with high angular
velocities (pk whereas the angular velocity of the platform and
the angular velocity \k of the procession remain sufficiently
low, so that only terms with the kinetic moments of the rotors
are kept in eq. (14). Then we have
u>-cfc = 0 ( fc = l , . . . , s ).

(9.12.17)
If the number of gyroscopes s > 3 and there are three non-
coplanar vectors
a; = 0,

(9.12.19)
that is the platform is stabilised in space. Equation (16) yields
s s
^2 HkCkXk =
m
° or ^2 Hkbk x ak\k + m° = 0
(9.12.20)
k=1 k=1

and expresses the ” equilibrium equation” of the platform, that


is the geometric 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.
a
i = bi = —iicosxi - i 2 s i n x i , ci = iisinxi - i2Cosxi, ' c2
13, a2 b2 = ii cosx2 + 13 = -iisinx2 + i3cosx2, > c3
a
= —i2, sinx2) =-iicosx3-i2sinx3- y
3 = 13,

b3 = —ii sinx3 + (9.12.2


1)
Condition (18) takes the form
cos (xi - X3)cosx2 ± 0. (9.12.22)
520 9. Dynamics of relative motion

It is not met at Xi = X3 i TT/2, i.e. when the rotation axes of the


first and third rotors become parallel to each other as well as
at X2 — TT/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

A\ (Xi + ^3) — (Ci — Ai) 12


W W COS 2xi + 2 (w2 - wi) sin

Hi (cu 1 sinxi — <^2 cosxi) = 0.


(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
(xi + ^3) + H1LU2
— ► (9.12.24)
M (X2 ~ ^2) — H2U3
= 0,
Within this approximation we reduce the three equations
obtained by projecting the vectorial relationship (16) on axes
i i , i 2, i 3 to the form
9.12 Gyroscopic platform 521

© Ty,*’ ©
ft'-* *ttt ■ © K-

FIGURE 9.7.
- H3 (u3 + x3) = m?, ' Q2W2 -
i?i (wi + Xi) = m$, Q3LO3 - H2 (9.12.25)
{ U2 - X 2 ) = ,y

Qi — ©? + C[ + C'2 + -A3 , Q2 — 0? + C's + Ai, Q3 — ©3 * + A2 •


(9.12.2
6)

Now we can cast the systems of equations (24) and (25) as

follows “■+=
^2 + -rlr“2 = d-m° i
(9.12.27)
MV2 V2
- Hi 1
“3 + ^s = gIra- J
Under the assumption that the projections of the resultant
moment m° on the platform axes are constant, the projections
of the angular velocity vary harmonically with the frequencies
H3 #1 H2
Ai A2 = A3 = (9.12.28)
VA2Q3 ’
which are very high since they are proportional to the angular
velocities of the rotors. We have
ou 1 = Mi sin (Ait 4- c r i ), ou2 = M 2 s i n ( A 2 £ + 02), ^3 = M 3 s i n ( A 3 £ +
03)
(9.12.29)
and, by virtue of eq. (25),

Xi = + M3 sin (A3t + cr3) - 4PpM2 cos ( A 2 £ + 02)


(9.12.30)
Hi VM
etc. The angles of rotation of the gimbals increase, however
small the moments 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 Mi, M 2, M3 produce
torques depending on the angles of rotation of the gimbals,
[43].
d2$ d2$ d2$
dx\ dx^dxi ' dxndx

«923
> d2<& a2#
dx\dxn dx2dxn ’' ’ dxl

(10.1.2)

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(2/1,... ,yn)(s=l,...,n). (10.1.3)

A. I. Lurie, Analytical Mechanics © Springer-


Verlag Berlin Heidelberg 2002
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 ^ is
introduced as follows
n
,Vn) = YhXkVk ,Xn), (10.1.4)
k=1
where all old variables are replaced in accordance with eq.
(3). Indeed, taking the partial derivative of ^ with respect to ys
and noticing that ^ depends on this variable, both explicitly
and in terms of the old variables, we obtain
3^ dxk dxk
dys ^ dys ^ dxk dys '
Xs + Vk X

since the sums cancel out by virtue of equalities (1). Thus


xs = — (s = 1, . . . , n ),

(10.1.5)
OVs
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
1n n
$ = 2 EE (10.1.6)
i=1 k=1

then its Hessian is the determinant |a| of the matrix a of the


coefficients of this form
($) = |a| (10.1.7)
Provided that a is a non-singular matrix, i.e. its determinant |a|
is not zero, then the system of linear equations
d<&
n
Vs dx„ Y^askXk (s = 1, . . . ,n) (10.1.8)
k=1
can be solved for the old variables. Denoting the inverse
matrix by b = a we find
n

xs — ^ ^ bskllk (5 = 1, . . . , n).

(10.1.9)
k=1
It follows from Euler’s theorem on homogeneous functions
an .. ai2 • flq n 2/
1 a &22 • • • ^2 n 2
i

21 /
2\a\
^nl <2n2 • • Vn
yi 2/2 • • • 2/n 0

(10.1.12)

An example of the application of Legendre’s transformation


is the generalised Castigliano theorem. Let us consider an equilibrium
configuration of a system with ideal constraints subject to
active forces of two types: potential forces prescribed by the
potential energy H(qi, ...,qn) and the applied loads F I , . . . , F A T - The
sum of the elementary work of these forces due to virtual
displacements of the system points from the equilibrium
position must 6'Wvanish
= ^2Fk-6rk-6U = J2 (Q*S ~ 1^) % = 0-
fc=1 s—1 ' '

Ql denoting the generalised forces of the loads. Provided that


the variations
6qs are independent of each other we obtain the equilibrium
equations
Q s = (s = 1
* Ws ’"-’n)

(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 qs of the system in the
equilibrium (10.1.14)
position under the above loads. The potential energy II is the
generat-
526 10. Canonical equations and Jacobi’s theorem

referred to as the complementary work. Relationships (5) are as


follows
qs =
dR
= ,n)
Ws -
(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. According
to eq. (11) the complementary work R is the associate
expression for the potential energy and Castigliano’s theorem
acquires the conventional formulation. The generalised
Castigliano theorem has a practical significance in those
cases when the complementary work can be explicitly
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)

10.1 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 system of n second order differential
equations for the generalised coordinates resolved for the
second derivatives qs 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 = rs and to introduce into consideration 2n
variables qs,rs which are independent variables determining the
behaviour of the system. The system of 2 n 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 differential equations
instead of n second order equations allows us to view all 2 n
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 q\ , . . . ,qn and the
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 qs
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, cp are solvable. This results in the relationships
(4.2.6)
qs = qa(pi,..- ,pn,qi,--- ,qn;t), (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 = ^2psqs — T = T (<71, . . . ,qn,pi,... ,pn;t), (10.2.6)
(10.2.3)
S=1
in which the generalised velocities are replaced by their
528 10. Canonical equations and Jacobi’s theorem

We obtain
^2 + Ps (%)*] = ^2 (
PsSqs
+ *
q 6ps
~ 7T~Sqs0qs~ JT~SpsPs+
s=l s=l \ °

or

Sqs +Ps [(<%*)* - <5<js] = 0.

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 d6 = 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
n
I dT +\ 6Qs = 0
'

Because the variables 8qs are independent we find

PS = -Q^-+QS (* = 1,... , ra).


(10.2.7)
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
or
n
H = ^2psqs-L = f + n. (10.2.9)
6=1
Function H depends on the generalised coordinates,
generalised momenta and time, whilst the generalised
velocities are removed by means of relationship (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 n.
dT _ dH
Qs = Q— Ps =
dT an dH

OPs dp s' dqs dqs dqs


An A\2 ... A\n Pi
A2i a22 ... A2n P2
1
m An i An 2 • •• A Pn
Pi P2 -.. Pn 0

We derive the system of canonical Hamiltonian equations r)H fW


Qs = Ps = -T- (s = l, ■■■,").
(10.2.10)
dps dqs
Variables qs and ps satisfying the system of canonical
Hamiltonian equations are called canonical
If the constraints are stationary then T is a quadratic form of
the generalised velocities and T is the associate expression
for the kinetic energy according to the previous section. As T
is denoted by T' in Sec. 4.2 we have under stationary
constraints
tf = T' + n,

(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'

(10.2.12)

As mentioned in Sec. 4.1, under non-stationary constraints,


dT2 <9T I
n
Ps — TP—I" IT ’
- = 2T2 + Ti,
8qs dqs s= 1
and it follows from eq. (9) that
n
H= - L = 2T2+T1- (T2 - TI + T0 - II) = T2 + (II - T 0 ),
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
530 10. Canonical equations and Jacobi’s theorem

are quadratic and linear in the momenta and a free term. The
corresponding expressions for T are given by formulae
(4.2.13) and (4.2.14). In eq. (4.2.13) T designates the
difference T2 — TQ 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 dH dH . dH \
dt dt dqs dps J
dH an an _ dH dH
dt dH
dqs dps dps dqs dt '
(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
H(qu... ,gn,pi,... ,pn) - ft,
(10.2.15)
which is termed the energy integral. For a system with
stationary constraints it states that the sum of the kinetic and
potential energies is constant.
If the coordinates # m + i , . . . ,qn are cyclic, then the
equalities dH
Pm+i = —z-------- (1 = 1 , . . . (10.2.16)
<JQm+1
which are consequences of the definition of Sec. 7.15 and the
Hamiltonian function (13), can be used to conclude that the
Pm+i=Pm+i (1 = 1,... ,n-m). (10.2.17)

In this case, the system of 2m first order


equations dH dH
(s = 1
4,
~W. ’’’‘'W. ............................ro) '
(ltl 218)

with the same number of unknown variables, which are the


positional coordinates and the corresponding momenta, is
under consideration. Assuming that this problem is solved
then the determination of the cyclic coordinates from the
equations
•o _ dH qm+i — dH
dPm-\-l d/? (/ = !, . . . , n — m) (10.2.19)
m +/
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

10.2 Explicit form of the canonical equations

In Sec. 7.8 we related the motion of a particle subject to


stationary constraints to the motion of the representative point
on the Riemannian manifold with the metric defined by the
square of the linear element
ds2 = Tdt2 = 1actf3dqadqt},
and introduced the velocity vector v (7.8.4) of this point with
the covariant components equal to the generalised velocities
qa. The relation between covariant and contravariant
components is given by the general rule (B.2.3) and in this
case takes the form
dT
a-aptf = = 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
representing the first set of the canonical equations. The
associate expression for the kinetic energy is given by
T' = ^aapa^p-fd^ps = ^ri^P-iPb ■ (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)
dV _ d (T + n) _ dH_
dpoc dpa dpa (10.3.4)
'
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
dU form
—Q a dqa'
Here, by analogy with eq. (7.8.8), we differentiate the velocity
vector v = para using rule (B.4.19) to obtain

Wa = Pa ~ 7<5 P } PPP6,
aj
and
dqa 1 /? (10.3.5)
Pa <27 P{3P6-
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
dT 1 da13'1
d^
By recalling Ricci’s theorem (B.5.8) and the rule of covariant
differentiation (B.5.6) we can write
VQa^7 = -77— + l }a"7 + { 7 \a^ = 0,a dq \ otp j
13 a
[ PJ

so that

-PfiPy= -UU}°" UM
137 +
1 da
2
c
dq P(3P7

and comparing with eq. (5) yields


. d (n + T) dH
Pa 77 77 ?
t'Qa dqa
as required. Thus, the explicit form of the canonical equations
is as follows
q' aa0pp, (10.3.6)

10.3 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, 1?, A, are given by expressions (7.18.6). We
obtain
Pr — mi?, P'd = mi?2$, p\ = mXR2 sin2 1?.

(10.4.1)
The Hamiltonian function will be
As pointed out above, the first set of the canonical equations is
nothing else than equations (1) resolved for the generalised
velocities
R = dpR P
dH dH Pd dH
A = dp\ Px (10.4.3)
R dp$ mR
2
’ mR2 sin
m
10.4 Examples 533

The second set of equations is


dU y
Pr = — dH
dR
Px '
sin2$, dR’
dH ► (10.4.4)
P _
P\ = o.
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',^ introduced in Sec. 2.10. The velocity vector of the centre
of inertia of the top is equal to
VC = v0 + (dn + ip sin x %z'c = vq + z'c $n' + ip sin ,
with vq denoting the velocity vector of the support point O. We

obtain
v
v
c=o + +z c sin2 + %zc ^vo • n' + ip sin $v0 • ,
and the kinetic energy is given by
T = (i)2 + ip2 sin2 d'j + Mz'c (^—dvo • n' + ip sin $v0 • +
(ft + ip cos d'j -\-^Mvo, (10.4.5)

where A and C designate the equatorial and the axial


moments of inertia
at point O.
The generalised momenta are
p^ = Aip sin2 d + MZ' VQ • n sin d + C + ip cos d^ cos
C

p& = Ad — Mz'cw o • n', p^ — C (^p + ip cos tf'j .


From these equations we obtain
ip = - 1 o . (Pxb ~ Pu cos d - Mzfcv0 • n sin d),
A sin d
#= + Mz’cv o • n ' ) , ► (10.4.6)
. Pep cos d , a .. . .
¥> = -£- — . ~o n (Ptl> - Pv c o s d - MZ VQ • C

nsintf).
534 10. Canonical equations and Jacobi’s theorem

FIGURE 10.1.
By means of eq. (2.113) we obtain the Hamiltonian function
1 2
H= -----------~—(pip — p<n cosi? — MZ'CVQ • nsini?) +
---------------(10.4.7)
2 A sin i?
2
2^ (p# + Mz'cv0 ■ n')2 + Yq+M9 (zccos# +
<o) “
Here Co (t) denotes the height of the support point. The terms

MgC0 - \Mvl
in the expression for H are immaterial.
While constructing the second set of the canonical
equations, we should consider that vectors n and n' are
functions of angles ip and $
n = ii cos^ + i2 sin^, n' = ( — i i simp + i2 cos ip) cosi? + i3 sinp, i'3
= — ( — i i simp + i2 cosip) sini? + i cosi?,
where i i , i 2, i 3 are unit vectors fixed3 in the space. Then we
obtain the following equalities
dn=0, dn' o di3 , 5ig
m -ncos#, — n, dip = n s i n p. (10.4.8)

Taking them into account, we have


dH w c/ dn
^v0 • — sin i? + P (v0 • n) cos i?
P0 dip _
= Mz'c (v0 • nsini?)* — Mz' WQ • nsinp,
C
10.4 Examples 535

where wg = vg stands for acceleration of the pole O. The second


equation
reduces to the form
dH cos, ■d 1 n(Pi{> - Pv $ - Mz'cv o • n s i n i ? ) - A s i n v
cos
.
Pip — dd
7)
. f n {p-d, — Pu, cos i? - Mz'cvo • n sin $) +
A sin v
{ dn' \
Mgz'c sin d + Mzfc ( vg • n^ cos d — $v0 • j .
The last terms can also be cast as follows
-Mz’c ( v0 • — p + v0 • -Q0&J = ~Mzc (vo • n')* + Mz'cw0 • n'.
Entering the quantities
P*4>=Ptl> ~ Mz'cv0 • n sin i?, p# = p# + Mz'cv0 • n',
we can recast the canonical equations
as follows 1) the first set
*=j^(p*-p«cos'(i)'
* = -7 Pi’ (10.4.9)
A1

. P<p cosP (p^-PvCosifj; ^


CA sin2 d
2) the second set
p^ = -Mz'cw0 • n s i n $ ,

PS = & ~ pv
cos cos d
~Pv)+ \
(10.4.10)
Mgz'c sin d + Mz'cw0 • n',
= 0.
i>%
We notice that due to eq. (8)
.Qd ,a
w g - n s i n d = —w0 • i3> w0 • n = w0 • hp

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
II0 = Mz'cWg • ig = MWg • Yc
into the expression for the potential energy. Equations (9) and
(10) are the canonical equations for the Hamiltonian function
(p*f - P<p cos I?) + ^jpf +
2
H*
2 A sin2 d
1
—p2 + Mgz'c cos $ 4- Mz'cw0 ■ i(3. (10.4.11)
2C
536 10. Canonical equations and

Jacobi’s theorem
where
wo • if3 = (xo sin ip — yo cos ip) sin d + zo cos d,

(10.4.12)
x0,yo,z0 denoting the projections of the acceleration of pole O

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 Ozf of the top. Denoting the rotation angle of
the flywheel by X, we set

a — ft = 0, 7=1

in eq. (4.12.3) which yields the formula for the kinetic energy
T= i A (j}2 + V?2 sin2 d^j + (fp + ip cos d^ +

^©3X2 + @3X (p + V’cosi?) •


(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 ©3 is the
moment of inertia of the flywheel about its axis of rotation.
Then we obtain
p# = Ad, = Aip sin2 d + C (fp + ip cos d^ cos d + ©3X cos d,
and
■a _ p± Pj> - Pf cos2 i?
A sin 1? ’
, Py-Px Iv-Pyeostf a ... C'Px ~ ®3Py (10.4.14)
* C ©3 A sin21? ’* 03 (C- ©3)'
These equations comprise the first set of the canonical
equations. The Hamiltonian function equals

Pl . (Pi>-P<e cosi9)2 A c
H
=2 A sin219 C ■ 03
P% + 2
Q-P X - PVPX2

+Qz c cos d,
f
(10.4.15)
where Q and zfc 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
£ V - = / 0 y, , Pv>=P<p, PX=0X■

(10.4.17)
The problem is reduced to quadratures, the distinction from
the problem of the heavy
_ top being
~ ©3 only in the determination of
x
~ e3(c-e3) Xo (10.4.18)

10.4 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(qi,... ,qn,pi,... ,pn) = u(p\q). (10.5.1)
v = v(q1,... ,qn,pi,... ,pn) = v(p\q).
Let us construct the
expression
du dv du dv du dv du dv
dqi dpi ''' dqn dpn
+ + dpi dqi dpn
referred to as the Poisson bracket of these functions dqand
n
denoted by
(«,«) = £ ( dqdu dv du dv \
k dpk dpk dqk ) ’ (10.5.2)
k=1 ^
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
(u, vi + v2) = (u, vi) + (tq v2).
538 10. Canonical equations and Jacobi’s theorem

which allows us in particular to write the Hamiltonian canonical


equations in the form

qs = (qs,H), ps = (ps,H), (s = l,...,n).


(10.5.6)

Finally, let us notice some values of the Poisson brackets

(Qs,Qt) = 0, (ps,Pt) = 0, (qs,Pt)


= 6st, (10.5.7)
with 6st denoting the Kronecker delta. These brackets are
known as the fundamental brackets.
If it is necessary one can use subscripts to indicate the
(U’V)q,p-
In what follows we use Jacobi’s identity, namely
(u ('v, w)) + (V, ('w, u)) + (w, (u, v)) = 0.
(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
(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
/ \ T~> (\ (7-, dm dw\
(”’w) = B w = E ( *9S +
B
)■

(10.5.10)
(«,«,) = A M = g + -4,,-.^
' dqk

where
= du dv du du
ddk ~7\ 5 -Hn+fc ~ 5 Afc = — , A _\-k — ^ (10.5.11)

n

OQk opk OQk opk


10.5 The Poisson brackets and the Lagrange brackets 539

Then
(u (v, w)) — (v, (u, w)) = A [B ( w ) ] - B [A (w)]

where dots denote the terms containing no second derivatives


of w. The expression under the double sum reverses the sign
under a change of summation indices s to k and fc 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,
(10.5.12)

This can be proved easily by means of the definition of the


Poisson bracket. Let us proceed to the expression
d^dpi dqn dpn dqidpi dqn dpn
du dv du dv dv du '" dv du
referred to as the Lagrange bracket and denoted by

(10.5.13)

Particularly, by analogy with eq. (7), we have


[qs,qk}= o, [qs,Pk]=Ssk, [Ps,Pk]= o, (s,k = 1 , . . . , n ) . (10.5.14)
These brackets are called the fundamental Lagrange brackets. When
needed we will use the independent variables q,p as
subscripts.
Let
^1 5 ^1 5 • • • 5 Vn (10.5.15)
(ui,m) ... (u1:un) («l,Ul) . . . (ui,vn)

(lin, U\) • * * (%) tin) (un,V l) ... (Un,Vn)


(vi,1ii) ... (vi ,u„) (Vl,V!) ... (vi,vn)

{vn,Ui) • • • {Vm ^n) (Un,Ui) • • • (%; ^n)


[[«>« [[«, V]]
]]
M [[«. «]]

(10.5.18)
see (A. 1.28).
By analogy, we can introduce the matrix A whose elements
are the La-
grange brackets
[,Ufc,'us],
These skew-symmetrical matrices can be written in the form
of 2 x 2 block matrices
((u,u)) ((•u,v))
(10.5.20)
((v,u)) ((v,v))

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
P' = -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 identity is proved
PA = AP = -P2n,
(10.5.22)
see eq. (A.2.39). The latter equation is equivalent to the four
((u, u)) [[u, u]]+ {(u, v)) [[v, u]] = -
E n, ' (10.5.23)
{(u, u)) [[u, v]]-I- ((u, v)) [[u, u]]=0,
( ( u, u)) [[u, ujj+ ((v, v)) [[u,«]] =0,
((V, u)) [[u, u]]+ ((v, v)) [[u, u]] = -En, J
10.6 Poisson’s theorem 541

For instance, an expanded form of the first equality is


n
^^ ^k) t^r] T Vk) \Pki ^r]} =
$sr•
(10.5.24)
k=1

10.5 Poisson’s theorem


Let us consider function (p of 2n variables q,p and time t. The
complete derivative constructed due to the Hamiltonian
canonical equations (2.10) equals
dp
uip v / dip dH dip dH
<p ■dt idn„ _ s dps dqdn„
dt \dqdn\ dn„
s dp s

which, using the Poisson brackets, can be recast


*=is+{'p-H)-
in the form (10.6.1
)
Thus, if (10.6.2
)
<p = (p(q\p',t)

is an integral of the equations of motion, i.e. ip is constant


throughout the motion, then

-£ + (<p,H) = 0. (10.6.3)

This equality is satisfied for all arguments g,p, t, i.e. it is an


identity.
Based upon Jacobi’s identity and relationship (5.12) it
presents no problem to prove Poisson’s theorem: if
<Pi(q\P',t) and

q>2{q\P^) (10.6.4)
are two integrals of the canonical
= (^1^2) equations, then the Poisson
(10.6.5)
has constant value c throughout the motion.
Indeed, according to eq. (3) we have the identities

%?± + (VuH) = 0, ^ + (</>2,tf) = 0 (10.6.6)


and it is necessary to prove
that
+ ((<Pi,<P2), ) = 0.
h

dt
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) - (<fi1, (tp2, H)) + ((</q, <p2), H)
= (<^2- (^l, H)) + {Vi > (H, V2)) + (H, (^2. V i ) ) =
which completes the proof.
The quantity rjj is either a new integral of the canonical
equations, or a function of known integrals, in particular,
integrals (4)
ip = V2 ) = "0(ci>c2)
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

dt a
' (10.6.8)
Continuing our reasoning we arrive at the
next integral
d2(p = ai
~dW
etc. Provided that ip = c is an integral containing no time t
explicitly, then expression (H,ip) 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 coordinates with one of integrals (3.5)
,Pm-,Pm+!>••• , / ? « ; * ) = c > ( 10. 6. 9)

we arrive at the identity


(<P,Pm+r) = = 0
(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
10.7 Canonical transformations 543

10.6 Canonical transformations


Let us consider two systems of variables, the old variables
<7i,.. • ,qn;Pi,--- ,Pn (10.7.1)
and the new variables
Qu... ,Pn, (10-7.2)
which are considered as functions of the old variables and
time t
Qs = Qs(<7 i , - . . ,qn;pi,--- ,pn',t) = Qs(q\p;t), 'j Ps =Ps(qi,---
,qn',Pi,--- ,Pn,t) =Ps(q\p;t) > (10.7.3)
(s = 1, . . . , n ). J
These relationships are assumed to be solvable for the old
variables, i.e.
qk = qk (Q\P;t), Pk=Pk(Q\P;t) (k = i,...,n). (10.7.4)
The transformation (3) is called canonical if the following
condition is met: the following expression
n n
'E'PkSqk-'^PiSQi (10.7.5)
fc = 1 1=1
is avariation of a certain function of the old and new
variables and time
t. Thetime 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


\r^ dqk _ p \ cn , % cn
/ jPk qq /j Pk z (10.7.7)
E p— dPi

The transformation will be canonical if the three conditions


are satisfied: firstly, in eq. (7) the derivative of the coefficient
of 8Qi with respect to Qi must be equal to the derivative of the
coefficient of SQi with respect to Qi
f dpk dqk d2qk \ _ dPj
fr'1\dQldQi dQldQiJ dQi
+Pk

y^ S'dp^dq^ d2qk \ dPt


^ V dQi dQi + Pk
dQidQi) dQi ’
0 En
— En 0

(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
{QhQi) =0, (Pl,Pi)= 0, (Qi,Pi)=8il ( i , J = l , . . . ,ra) (10.7.12)
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
(u’v)qP = J2 du dv du dv
k=1 dqk dpk dpk dqk J

the partial derivatives with respect to the old variables by the
partial derivatives with respect to the new ones by the
formulae
dQr d dPr d
—d= Y
dqk ^ —( \ dqk dQr dqk dPr J ’
j^fdQs 8
d = . 0PS d
+■
dpk V dpk dQs dpk dPs J ’

inserting into the bracket we obtain, after


summation du dv . . du dv _ „.
(u,v) qp over
EE
k,
( Qr,Qs)„p + •STT'oTr (Qr,Ps)ap +
dQr dQs dQr dPs
s=l r=1 (.P ,Q ) +— — (P ,P
r r s
du dv K r,Vs) -r QP QP K T, s qp
s
gp (10.7.14)
dPr dQ.
In the case of the canonical transformation, that is under
condition (12) we can write

(u,v),
du dv du dv
qp EE
rj \dQr dPs dPr dQs

du dv du dv (u,v) QP ■
=E —^ \ dQr dPr dPr dQr
546 10. Canonical equations and Jacobi’s theorem

which completes the proof. The invariance of the Lagrange


bracket under the canonical transformation, i.e.
M m = [uMop (10-7-15)
is proved by analogy.

10.7 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 asmuch as there exist relationships (7.3) and (7.4)
it suffices to consider it to be dependent on 2 n arguments and
time £, among them n old and n new arguments. Thus we have
four types of generating function
Vi(q,Q;t), V2(q,P;t), V3 (p, Q; t), V4(p,P-,t). (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 5
J2 (p* « - pi8Qi) = 6 V i f a - G ; * ). ( 10. 8. 2)
2=1

n
T (Pi^Qi + Qi&Pi) = Vi (q, Q; t) + ^ 6V2(q,P;t), (10.8.3)
& QiPi
2=1 . 2=1 .

-Y,(qiSpi + Pi6Qi)=6 Vi (q, Q;t) - ^ qiPi SV3(p,Q-,t),


2=1 2=1 (10.8.4)

n n
^2 (qiSpi - QiSPi) = 6 Vl (Q’) Q\ t) ^ ] QiPi ^ ^ QiPi
2=1 2=1 2=1
= SV4(p,P;t). (10.8.5)
From these equations we obtain the following systems of
canonical transformations
dV2 n dV2 Qi~ dp
Pi = dqi' ,n), (10.8.7)

dVj
Qi = .. ,n) (10.8.8)
dpi ’

dV4
Q
Qi = dpi ’ ‘=w ( * = 1 - . ,n). (10.8.9)

Let us consider, for instance, transformation (6). Taking a function


Vi (qi,-.. ,qn;pi,--. ,Pn,t)

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
<92V d2Vx
i dQidqn
dQidqi ^0. (10.8.10)
d2Vx
dQndq
2
d Vy
dQnd n
qi
This is the requirement imposed on the generating function V\.
Expressions for the new variables Qs in terms of the old ones q
%,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
d2V2 d2V2
dPidqi dPxdqn
d2V2
dPndq 7^ 0, (10.8.11)
d2V2
i dPndq
n

and the procedure for constructing formulae for


transformation (7.3) reduces to determining the new variables
Ps 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
d2V} dpi dPs d2V2 dpi _ dQs
dqidQs ~ dQs ~ dqi ’ dqidPs ~ dPs dqi
d2V3 _ dqi dPs d2V,1 dqt _ dQs
dpidQs dQs ~ dpi’ dpidPs ~ dPs dpi

(10.8.15)

(10.8.16)

Using these equalities let us find the Jacobian of the


canonical transformation [16]

Ql; ■ ■ • ? Qn] Pi ? Q\P


D • • • ? Pfi =D
<71,... ,<?n;p q\p J
By virtue of the property of Jacobians we can write

Q\P Q\P\ D(q\p) q\Q (10.8.17)


D =D
\ Q\ P J ' D{q\Q)'
10.8 Generating functions 549

On the other hand,

By analogy we also obtain


q\p Pi,-- - ,Pn
D q\ Ql, - - - ,
Q Qn
and the first relationship in eq. (15) takes the form
Q\ (10.8.18)
D P
q\Q
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
Q\ = l. (10.8.19)
D P
q\p
The consequence of this statement is Liouville’s theorem on the
invariance of the integral
H dqi • • • dqndpi ... 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
S-Kw) dQi. ..dQndPi. ..dPn.
dqi.. ■ dqndpi... dpn

Here the integration on the right hand side is carried out over
volume V resulting from v under this transformation. For the
canonical transformation
550 10. Canonical equations and Jacobi’s theorem

(19), we have
II II dQi... dQndPi... dPn.
dqi. . . dqndpi ...dpn 2 n
(10.8.20)

10.9 Invariance of the canonical transformations

Let us recall that the variables satisfying the system of the


canonical equations, 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,PS related to the canonical variables qs,ps by a canonical
transformation are also canonical, that is, they satisfy the system of
canonical Hamilton’s equations
(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+j£,

(10.9.2)
Vi being the generating function of the canonical
K = H. (10.9.3)
Let us consider the transformation with generating function
V\ (q, Q; t).
The variables qs, Qs are considered as independent,
and applying this transformation we have
dQ
s
= 0. (10.9.4)
dt
On the other hand, we obtain from the second set of formulae (8.6)
_ d2Vx
dPs
(10.9.5)
dt dtdQs'
Let us consider now QS,PS as functions of time and the old
canonical variables prescribed by eq. (8.6). Then, based upon
relationships (6.1), (4) and ( 5 ) , we can write down the
equalities d2Vi
Qs = (Qs, H)qp, dQsdt+ (Ps,H)qp, (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
dH
Qs — (Qs5 H)QP = Qp > (10.9.7)
.J-fW+H
dQs V dt

What remains is to notice that


dPs dPs \dt )

since V) does not depend on Ps. Denoting


dVi
K=H+ dt ’

we arrive at the system of the canonical equations (1)


Q =**L ok
dp, ’ p, = - dQs (s = 1, . . . ,n).

which is required. Nothing would be changed in this derivation


if the generating function V3 was used.
In the case of the generating function V2 (or V4) we should
take
3Qs d2V2 dPs
dt dPsdt' dt
instead of eqs. (4) and (5). This results in

dQs d2Vo
Q
s dt + (Qs,H)q dPsdt 'Q dP dt
P
dH s +H
P dQs dQs d dV
2
since Qs does not appear in expression V2 (q,P;t). Taking

dV
K = dt + H,
2

we arrive at the canonical equations.


Thus it is proved that the canonical equations are retained
under the canonical transformation of variables.
dh dh
dqi ‘ ‘ ' dqn
dfn " ’ dfn
dqi ' ’ ’ dqn

(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)
n s_ dV2 _ f (
V — Qp — Js \Ql ? • ,qn) ( s = 1 , . . . , n ) (10.10.3)
• •

describes the point transformation of the generalised


coordinates. Expressions 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
p ®fs /• _ 1 \
Pi
2^Ps f)n. — 1? • • • ,n). (10.10.4)
S=1
' dq,

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, A.
The kinetic and potential energies are given by

T= (^z2 +f2 + r2A2^ , II = / ^\Jr2 + z2^ ,


and the Hamiltonian function is
H
= 2“ (p*+^ + ^) + /(v/r2 + * 2 ).

(10.10.5)
In order to transform to spherical coordinates we should
V2 = Pi \J z2 + r2 + P2 arctan - + (10.10.6)
P3A,
z
10.10 Examples of canonical transformations 553

which can be viewed as the generating function of the


canonical transformation since condition (2) is met. We obtain
by means of eq. (3)
Qi = y/z2 + r = i?, Q2 = arctan - = $, Qs = A, (10.10.7)
2
z
R, $, A denoting the spherical coordinates. By virtue of
formulae (4) we find
Z T T Z
Vz = Pl
~R~ P<2~&' Pr = + P2
H’ Px = Pa’
and the Hamiltonian function, in terms of the new variables, is
obtained by substitution of these expressions into eq. (5)

" = 2k(F?+ip|+>?)+/<R)
=
(Pl! h ih + /i'2,-:nr ) + 'i [R'
(10.10.8)
where PR,P^,P\ are the new momenta corresponding to the
10.10.1 Second, example
Let us consider the transformation
OF
Qs = qs, Ps=Ps~ ,
(10.10.9)
OQs
in which F is a function of qi,...,qn and t. To prove whether this
transformation
S
is canonical, let us construct expression (7.7)
’ r
-f \ Finn Finn Finn Finn /

^ V dqk dpk dpk dqk )


d8rk+8sk
dF
22
F d2F d2F d2F
~1^\ f)a,rta + ~ ~ = o.
k=1 dqkd dqkdqrJ dqrdqs dqsdq,
By analogy q,
(QsiQr') = (qs,qr) = 0
and
0Qs,Pr) = J2 OQsdPr dQs dPr\
/ J VskVrk — Os
fe=1 \ 0qk dpk dpk dqk/ fe=i

This transformation is seen to be canonical. The generating


function can be taken in the form
v
2 = Y1 qsPs + P(qi>-- - ,qn,t). (10.10.10)
s=1
554 10. Canonical equations and Jacobi’s theorem

10.10 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 1 n
^ Qy'
T = -MVQ + -u> • 0° • u> + M (v0 x u>) • r'c + Mv0 • V] +
Z Z CfQs
s=1

N n
dr'-
V -^772^' X ’
_hTr
(10.11.1)
i=l s=1 0(*s

where r' ( # i , . . . ,qn) denotes the position vector of the particles


with respect to the pole O of the coordinate basis fixed in the
carrying body. Its motion, described by the velocity VQ of the
pole O and the angular velocity vector u;, is assumed to be
given. The position vector of the centre of inertia referring to
the pole O is denoted as y'c. Finally, Tr is the kinetic energy of
the relative motion which is a quadratic form of the
generaliseddTvelocities.
dT The expressions for the momenta are
- - -= —-r + U) £ miri , +dv\ Mv
0
drc
(10.11.2)
Ps
dqs dqs i—
x

dq

dq
1 s s

and the Hamiltonian function, due to eq. (2.13) is equal to H =

Tr + n - To = T; + n - ■ e° • a> + M (u> x v0) • r'c,


since the term —MVQ in (1) depends only on time and can be
(10.11.3)
omitted for this reason. Replacing in the expression for Tr the
generalised velocities in terms of the momenta we obtain T’r.
Let us enter matrix B — A~1, which is the inverse matrix for the
matrix A of the coefficients of the quadratic form Tr. Then we
have, due to eq. (2),
n n N
T
i ( / \
r= x

s=l k— 1 \ i=1 S
/
(Pk-u-J^nur'iX (10.11.4)

Here the quantities


dr' 8
Ps=ps-MvQ-—£-=ps-—Mv(i-Y'c (s = 1,... ,n) (10.11.5)
Oqs OQs
are introduced. Adopting that

Qs=qs (s= 1,... ,n), (10.11.6)


10.11 Canonical equations of the relative motion 555

then the transformation relating the new variables P s, Qs to the


old ones ps,qs will be canonical of the type (10.9) with

F = Mv0 • r'c = M (v0ix'c + v02y'c + ^03zc) ■

(10.11.7)
The partial derivative of F with respect to time is equal to
dF *
— = M {v<nx'c + V02y'c + v03z'c) = M v0 -r^, (10.11.8)

and the Hamiltonian function K for the system of the canonical


K = T'r + n + n + Mr'c ■ (v0 + u x v 0 ) = r; + n + n“ + n°.
w

(10.11.
9)
Here, according to notation (9.2.10) and (9.2.9), n w and n°
denote the potential energy of the centrifugal forces and the
inertial forces of the pure translation, respectively. The value
of T’r is given by eq. ( 4 ).
The system of the canonical equations of the relative motion
is written in the4sform
=
dpr Ps = ~Ws = (10-1L1°)

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 and n° 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
p{;) =
di (* = 1
>•••'")•
(10.11.11)
The associate expression for the kinetic energy in terms of
these quantities can be represented by the quadratic form
and, by comparison with eq. ( 4 ) ,
yields
C) + u ■ ^2 mi*i X (s = l,...,n), (10.11.13)
P. N dq.
dK dr; dn dn^ dn
dqs dqs dqs °dqs

1
1

(10.11.16)

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 dr'
E miTi x dqs ’
i~ 1
depend on qs, so that by virtue of eqs. ( 4 ), (13) and the first set
of equations in (10), we have
_ 1^
dT'r ^ 9Bik (r) (r)
a** ~ " ~
Pi n
1=1
k=1
nn (r) dr(
(dr[ d2r'
1 = 1 k=l dq s dq k
^ \ sdqk
dq
d*Ti /dr' c>r' , a2A
w 2=1 m X +r
-dq's E E d 7 ^ f ^Vd^s X
’ df/fe dqsdq qk. (10.11.17)
k=1 i=l k
10.12 Canonical transformation and the process of motion 557

The first term represents the derivative of T'r 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)
2 = becomes
1
p ; = q\
{ } dr'. c)
dqs dqk dqs dqs n + ir + n0)
or
"(r) = ^ (n + n- + n°) + cfa + Q(scs=
or
1, . . . , n ),
P'S ' =
(10.11.18)
whilst the first set, due to eq. (12), is given by the equalities
dV
Qs = -777 (5 = 1,... ,n).
(10.11.19)
dpi J

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 transformation of the canonical equations of the
absolute motion, is an intermediate form between the latter
and the equations of the relative motion. Because of the terms
IT*' and II0 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

10.11 Canonical transformation and the process of


motion
It is assumed that the general solution of the system of
canonical equations
• dH ■ 9H . 1 ,
, =
' 5? p‘ = ~w. (s = l " " - B ) f1012-1)
is known. This solution yields expressions for the generalised
coordinates
and momenta as functions of time and 2 n arbitrary constants
denoted by
otk and f3k
(10.12.2)
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
(<?!,••• ,Pn) ^
(oq,... , OLn 5 P\5 • • • Qi Pn) (10.12.3)
D

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
Qts --- &S (L Ql 5 • • • 5 Qn 5 Pi 5 • • • 5 Pn) 1
Ps = P s ( t , q i , - - . ,Pn), (s = 1,... ,n). (10.12.4)
These expressions provide us with all 2 n integrals of the
canonical system
of equations (1). Hence, according to Poisson’s theorem any
of the Poisson
brackets (a8,ak), ( a „ / J f e ) , (P8,Pk)
(10.12.5)
is either constant (in particular equal to zero identically) or is
expressed in terms of the same constants a^,/^; otherwise the
brackets (5) would contain the (2 n + 1) — th independent
integral which is impossible.
Assume now that we took not an arbitrary system of 2 n
independent integrals of the canonical systems of equations
(1) but its total Cauchy’s integral. This implies that constants ak
and (3k are(Qs)t=tdefined
0 ~
as
(Ps)t=tthe
0
values of the generalised
— Ps (s — 1,... ,n).

Then, recalling the definition (5.13) of the Lagrange’s bracket,


we obtain, for example,
{as Pk
^ / dqr dpr dqr dpd r \
' '~ ^das)

and thus
n
[aS5 Pk\t = t0 / dar d(3r dar ^ &rs&rk ~
V<9as d(3r &sk' r= 1
d(3k d(3k
By analogy we find
[o!s, CVk]t_to 0, IPs’) Pk\t=to
Based on these equations and relationships (5.22) (or (5.24))
we obtain similar expressions using Poisson’s brackets
(asi Pk)t-tQ = (as, ak)t=to = (Psi Pk)t~to ~
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,
(<*«, (3k) = S8k, (as,ak) = 0, (/3S, (3k) = 0,
(10.12.6)
which means that relationships (2) representing the Cauchy
integral of
the canonical equations are formulae for the canonical
transformation of
variables ak,f3k 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 (10.12.7)
has the
generating
function
is of the integral
the Cauchy canonicalof transformation
the system ofhas,
theis canonical
of great
equations with the following Hamiltonian function
H = I(AV +p2) .
(10.12.8)
Equations (7) describes the motion of an oscillator. Let us
prove, by calculating Lagrange’s bracket, that equations (7)
presentrlaa,a dq
canonical
dp transformation.
dq We have
’0] = -^m cos2 A (t — to) + sin2 (t — to) = 1,
dp
which is required since d/3 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 i, y have the form
X = x0 + x0 (t - t0), y = yo + yo(t- t0) - ^g(t - t0)2 ,
(10.12.9)
x = ±0, y = y-g(t-t0).
These equalities are Cauchy’s integrals of the system of
canonical equations with the following Hamiltonian function
H
= \ (±2 + v2) +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
[x0,yo], [zo,i0], [x0,yo],
[yo, io], [2/0,2/o],
[io,yo] •
A simple calculation shows that all conditions (7.8) are fulfilled.
For instance,

r , dx
|l0 wl =
dx dx dx dy dy dy dy
’ a^ dy0 dx0 dx0 dy0 dy0 dx0
and so on.

10.12 Jacobi’s theorem

It is proved in Sec. 10.9 that under the canonical transformation


Qs=Qs(q\p;t), Ps = Ps(q\p;t), (s = l,...,n) (10.13.1)

the new variables Ps, Qs satisfy the system of Hamiltonian


equations
Q = 2IL p _
dK
dPs' s dQ, (s = 1 , . . . , n ) , (10.13.2)

obtained by means of the Hamiltonian function

(10.13.

3)
9H 9H
Here *
denotes (U 1
the(fcHamiltonian \
function for the initial
H Qk = ^~, Pk — ~~~p\ = 1,... , n) dpk
dqk
canonical
and V is variables
the generating function. Clearly, K should be
expressed in terms of the new canonical variables Ps, Qs •
Thus, given a generating function V, we reduce the problem
of integration 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 ), Ps and Qs are constants
Qs ^5) Ps= Ps (s = 1,... ,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
(3S, then its determination, due to eq. (3), reduces to the
solution of the Jacobi-Hamilton equation
dV ( dV dV \
— + H[qu... =0,
(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
dV
The complete integral of the Jacobi-Hamilton equation (6) has
n constants Additionally, it must satisfy condition (8.11) which
has the form
d2V d2V
dpidqi d(3i
d2V dqn 7^0 (10.13.8)
d(3nd
qi d2V
d(3ndq
under the adopted notation. In particular, if at least one of the
n constants appears in V additively, i.e. if

... 1 qni /^11 • • • 1 @n— 11 "b


V (q 11
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)
dV=at dV , , , ,
Wt - di = n (10 13 9)
' '
The first of these systems is solvable for <71 , . . . , qn by means
of eq. (8). The solution yields expressions for q\, . . . ,qn in terms
of t and 2n constants as,f3s. Inserting these expressions into the
second system, we find the momenta as functions of the
same values £, as,(3s
Qs qs ... , Oin; , /3n),
ps =ps(t1a 1,... ,/3n) (s = l,... ,n). (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. (10.13.11)
Let us notice that the general integral of the canonical
system (10),
obtained by Jacobi’s theorem, is a canonical transformation of
variables
as, (3S to variables qs, ps carried out with the help of the
generating function
V. Any other general integral of the form

Qs Qs (L , • • • , Cj7,5 D\ , • • • 5 Dyf) ,
Ps — Ps (L 5 • • • 5 Cn 5 -^1 5 • • • 5 TJqf) (s =
1, . . . 5^)5
Ck,Dkbeing 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
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 (qi,... , / 3 „ _ i ;h) (10.13.14)
and satisfying the condition
d2W d2W
diiydqi dPi
2
dW dqn
7^0. (10.13.15)
d(3n-id<h d2W
d2W dpn-\Pqn
dhdqi d2W
dhdqn
We can construct now equations (9) which take the form
dW
——
dP= ak (k = 1, . . . ,n - 1 ). (10.13.16)
k

dW
dh t to? (10.13.17)

Vk~—{k (k- 1
Pk
~
dqk ~1' ,n). (10.13.18)

The constants h and £o play the part of f3n and an.


Equations (16) are n — 1 dependences between the
generalised coordinates #i, and do not contain time. They
produce a set of all trajectories 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 2
d 2W dW
d^dqi d(3idqn-i
dW2 ~f~ o*
d2W
dfi^dqx "■ d&n_xdqn-1
Equations (16) yield then the parametric representation of the
trajectories
qs =qs{qn',oi1,... , a n _ i ; / ? ! , . . . , / 3 n _ 1 , / i ) (s = 1 , . . . , n - l ) ,
(10.13.1
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~ t0',ai, • • • ... ,Pn-i,h) (s = 1,... ,n) :
(10.13.2
0)
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-t0;a1,... ... ,Pn_i,h) (s = 1, . . . ,n).
(10.13.2
1)
Relationships (20) and (21) represent the canonical
v = -ht + W(qlt... ,qn, Pi,-- • ,Pn-i,h) (10.13.22)
of the type V"2 of coordinates qi,... , qn and new momenta fl1; . .
■ ,/3n_1,h.
Provided that <jrm+i, are cyclic coordinates, the Hamilton-
Jacobi
dW dW (10.13.23)
H
(qu...,qm, Qqi,...,dqm,...,dqn) h,

and its solution should be sought in the form


W = 0mqm+1 + ... + + W* (qi,... , qm) (10.13.24)

Substitution into eq. (23) leads to the Hamilton-Jacobi
equation for function W*
dW* dW* Pn-l = h- (10.13.25)
H [qi,... ,qm,
dqi ’' " ’ dq„

The complete integral of this equation contains, apart from n -


m constants (3m,... , /?„_!, h, other m— 1 constants Pi,. ■ ■ , Pm-i -
The following condition
d2W* d2W*
dPidqi dp 1 dqm
d2W* d2W* 7^0 (10.13.26)
dPm- dPm_idqm
idqi d2W*
d2W*
dhdqi dhdqm
V3 (p\Q\t) = Vs (pi. > * • • ) Pn? Oi, . . . , ^) • (10.13.29)
Then, by virtue of eq.
(8.8)
dV
(10.13.30)
dV
a (s = 1 , . . . , n ) ,
Qs = Ps = dps UOl-S
and the Hamilton-Jacobi equation will take the
form
(10.13.31)
dt \ dpi ’ ’ * * ’ dV
a ~-Z—,Pl,... ,Pn,t =0. uPn J

Here the generalised coordinates among the arguments of


the Hamiltonian function H are replaced by the partial
derivatives of the required generating 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 system 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
f31, . . . , /3n_1, ft, then the relationship
W{q1,... ,qnA, . . . h) =7,
(10.13.32)
7 being a parameter, can be deemed as an equation for the
family of hypersurfaces of the Riemannian manifold Rn. Notice
that superscripts are used in the latter equation since tensor
566 10. Canonical equations and Jacobi’s theorem

where the variations 6qs are the contravariant components of


the infinitesimal displacement vector <Sr on this surface, so
that <Sr = rs6qs, rs are the basis vectors and 6qs = rs • <Sr.
Recalling equalities (18) we can write eq. (33) in the form
dW s
——r • 6r = psrs Sr — 0.

(10.13.34)
oqs
Thus, the momentum vector (or the velocity vector of the
representative point)
w=ps r 5,
(10.13.35)
p = grad W.

(10.13.36)
This equality becomes especially illustrative in the case of
p = rav = grad W.

(10.13.37)

10.13 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 (10.14.2)
form
1 n* / \2
- y ^ A f c ( 9i , . . . , g n ) ( — ) + H ( q 1 , . . . ,qn) = h. (10.14.1)
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 r and cp. The expression for the Hamiltonian
function for a particle of unit mass is
f_ (10.14.3)
M_
r
According to eq. (13.24), the complete integral of the
Hamilton-Jacobi equation
dw\2 2i /aw\ fM
dr ) r \ dip )
r
should be sought for in the form
W = 0v<p + R{r), (10.14.4)
where R (r) is defined by the following differential

equation (10.14.5)

We obtain then that \R' (r)]2 = 2h+ — — n = fM.

W = 0v'p±J )j2h+^- &dr. (10.14.6)

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
r
dW dr
J

r (10.14.7)
dW dr
dh =±/ = t — to.

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

&w_y 1 fdW \ 8RJ + 1 fdWY E h. (10.14.8)


R2\dd ) + R2 sin2 d V OX ) R

Looking for a solution in the form of eq. (2)


W = (3X\ + Wi (R) + W2 (d),
we arrive at the equality

d W i Y +2h+
f t I 2/i
dw2y , pi
R ~dR R dd + s i n2$ „Q ’

which can hold only if both sides are constants. Denoting this
constant by /?$ we obtain two equations

and the complete integral of the Jacobi-Hamilton equation is given by

ji 2
W = (3X\ + J )J2h + ^-§dR + J. I sin dd. (10.14.9)
Fi d

From the latter we obtain the equation for the


trajectory
-PxJ- dd a\,
Pi
sin d\ Pi -
2
sin2? > (10.14.10)
dR ?
J W2/, f-g + I dd = OL$
+ Pl
IPl--
sin2 d
and the equation for time

dR
= t- t0. (10.14.11)
/ 2/l +
l“P
10.14 Separability of variables in the Jacobi-Hamilton equation 569

lO.l^.S 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\ and O2 are taken to
lie on axis Oz of the cylindrical coordinate system r, <p, z the
distance O1O2 being denoted by 2c. Let the position of the
moving particle M in the meridional plane cp — const be
described by the generalised coordinates g i , g 2 , which are
related to the Cartesian coordinates r, 2 in this plane by

where 1 < q\ < 00, — 1 < g 2 < 1. The curves q\ — const, q2 =


const are the confocal ellipses and hyperbolas

having the foci at the centres of attraction z = ±c. The sum rq


+ r2 and the difference rq — r 2 of the distances from the
attracting centres depend only on qi and g2, 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,
r\ 2 = (z± c)2 + r2 = c2 (qx ± q2)2 ■

or
ri = c (qi +q2), r2 = c (qi - q2).
The expression for the potential energy has the form

It is easy to calculate the kinetic energy


T= I(i2+r2+r2^)

The Hamiltonian function becomes

2cz qt~q2 q{-q2 Wi - iM1


, 21 [(/1 + h) qi + (/2 - /1) •
(10.14.12)
570 10. Canonical equations and Jacobi’s theorem

Looking for the complete integral of the Hamilton-Jacobi


equations in the form
W = 0v<p + W* (<?i,q2),
we arrive at the equation

(«? -1) dW dW* 0 2 02


* +
dQ2 J ' q - 1 'v___|
2
1 - ql
<P

dqi
2c [(/i + f2) qi + (/2 - /i) 92] = 2he2 (q\ - ql) ,
from which one can conclude that the variables q\ and <72 are
separable. Indeed, assuming
w* (qi,q2) = W1(q1) + W2(q2),
yields the relationship

(«? - !) (l£~) + 'J~[ ~ 2c(h + &qi ~ 2hc2Q2i

dW2\ 0i
dq2 ) + 1 - q\ ■ 2c(/2 - fi) 92 + 2hc2ql

Equating each side to constant — /?, we obtain


«-1) ( dWx 0 V + 2c (/1 + / ) 91 +
0-----2 2
dqi 2he Qi
2
ql
(1 - 4) dW 0lQ1 t
2 = 0- + 2c (/2 - fi)q2- 2hc ql.
2
dq2 ) ^ l ~qf
Expression W is now represented in
W = 0v<p+ f
the form 1
+f j-prr,1 (10.14.13)
J Qi ~ J Q2 ~

where
Fi(qi) = {q21-l)[2hc2q2+2c(f1+f2)q1-0\-0l,
F2 (q2) = (l - ql) [-2hc2ql + 2c (f2 - /i) q2 + 0] - /?2 .
The equations for the trajectory and the time take the form
dq2
[_________dJl__________[
<P~0L ’J (q2-l)y/FM PlpJ (i -QDVFM
OLtr
_ f dq i f dq2
J VK{qi) J ' = 2 a,
\JF2
, f q2dqi _ c2 f (q2)
J VFM J = t-t 0.
q2dq2
\/F2 (q2) (10.14.14)
dpi 9pn d2W d2W
9(31 " Pndp1 ’ dfiidqn
= Pi••-Pn
dpi d2W d2W
9pn '• Pn
9h dhdqi ’ ' 9hdqn

(10.14.1
7)
is not zero. This system gives

As = ^ (s = l,...,n), (10.14.18)

where Ans denotes the algebraic adjunct of the element of the


n — th row
and the s — th column in the determinant (17). Inserting these
11
=

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)
572 10. Canonical equations and Jacobi’s theorem

only variables i.e. the following equalities

Pk^rr = <Psk (Qk) (s = l,... ,n-l),


P
Qp '
(10.14.20)
Pk-j^ = Vnk( k) (k = l,...,n) q

must hold. The n2 function (psk (qk) are arbitrary; the only
requirement is that the determinant obtained by substituting
(20) into (17) does not vanish, i.e.
Vhl(gi) ••• Pi n(qn)
7^0. (10.14.21)
<Pnl(<h) ■■■ VnniQn)
The consequence of eqs. (2) and (15) is the
formulae
_ fdw\
Pk \dqk ) (10.14.22)

On the other hand, equalities (20) allows us to write


n—1
PI = 2Y1 PsPsk (Qk) + 2hipnk ( qk ) + 2 ^k ( qk ) , (10.14.23)
S= 1

where ipk (qk) are n arbitrary functions. We


n n * n— 1
= E - = T,j\2T.>
obtain
w PsVsk (Qk) + 2htpnk (qk) + 2ipk
(qk)dqk.
k =1 =1 k J
\ S=1

(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
n—1
*nk ^2 ^fc) + h(fink (Qk) + V’fc (Qk)
k=1 .5=1
The determinant property
^Psk (Qk) =
k=l

yields the following equality


n
=~J^A (Qk) —= - J2 Ak^k (10.14.25)
k=1 k=1
10.14 Separability of variables in the Jacobi-Hamilton equation 573

We come to Stackers theorem: taking n(n + 1) functions (psk


and ^k(qk) we can construct the Hamilton-Jacobi equation in
(qk)
form (1) in which the coefficients As and the potential energy n
are given by formulae (18) and (25), Ans being the algebraic
adjuncts of the elements of the last row of determinant (21).
Formula (24) gives the complete integral of this 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 (psk (qk) and ipk(qk) such that the coefficients As and
the potential energy n 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
nr Vsk (Qk)dqk
dW
dP OL s-£/
n—l 1/2
s k=lJ 2 PrVrk (Qk) + 2htpnk (qk) + 2ipk (qk)
r— 1
[
dW u p .
dh -*o = £/ ---------
(s =- 1, . . . , n - 1 ) ,
k=iJ r ^ 1 1/2 ‘
2 PrVrk (Qk) + 2/i</>nfc (qk) + 2V’fc (qk)
r= 1
(10.14.26)

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 cpsk (Qk) and 'pk(qk) in terms of given As and n. According
to eqs. (16) and
<Pn (qi) + A2(f12 (#2) + • • • + A^in (Qn) — 0, '
^l^n-1,1 (Ql) + ^2^n-l,2 (^2) + • • • + An(fn_l n (qn) = 0,
*
A
l<Pnl (?l) + A2^n2 (&) + • • • + AfiVun fan) =1>
(10.14.27)

is fulfilled for given As. This problem can be solved only if


functions 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.
Oi — d2 0 ... 0 0
0 Cl2 —a3 ... 0 0

0 0 0 ... tin—1
a\b\ a2b2 <2363 ... Q"n—lbn
—l

The row of functions ( < Z fc ) j obtained by means of condition


(25), is
^k — akCk •
In formula (24) we have now
n~ 1 ( Plal (&=1),
^PrVrkiQk) = l (~Pk- 1 + P k ) ( k = 2 , . . ., n — 1 ) ,
r=l ^ -Pn-iOn (k = Tl) .
To simplify the notation we put

Pi ~ Pli P2 P2 Ph * • • j Pn— 1 — Pn-1 Pn-2i Pn = ~Pn-!,


10.14 Separability of variables in the Jacobi-Hamilton equation 575

so that
n— 1
^ ^ Pr^rk ((Zfc) Pk^k 1, . . . , 7lj ,
r=l

under the condition

XX = °-
(10.14.31)
fc = l
The complete integral of the Jacobi-Hamilton equation is
written as follows
W = k=I
J
2 akckdqk, (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

h = q\, b2 = -ql, ai = a2 = ,2 ’
/2-/1
fi + /2
c i =-----------------9i, c2 -92-
We obtain then that

W = cj + 2hq\ + 2L±A9l ^L_ +

c ( J2/32 - 2 hql + 2 ^ ■2,


JV c v 1 — q2
where (31 + P2 = 0- 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 (9? - !) + <P12 (! - «£) + <P13 (~Z12~ 71 —+ 0
\9i ~ 92 ,
<^21 (9? - !) + V>22 (1 - 92) + ^23 ( -2^—T
+ T~2 0
,
fzi (<?1 ~ !) + ^32 (l - ?!) + <^33 ( ~2 7
+1 12
\9i — 1 1
~ Q2
576 10. Canonical equations and Jacobi’s theorem

The functions ips3 can not depend on the third coordinate <p,
hence they are constant. Letting <^31 = (p32 = — L ^33 = 0 we
have
Vsl ( 9? - 1) + ^Ps2 (! - <&) (* = 1.2)
l-9l

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
<Pu (9? - !) - T2~“T = °. ¥>12 (! - 92) - = °.
9i - J- 1
“ 92
^21 (91 - 1) - “2--71
= !. ^22 (1 - <£) - 7-1 —- --2 = -!•
Qi ~ </2
The matrix of functions (</&) is given by

(^-l)2 (1 - 922)2
9i 92 _j
(9?-l)2 (l-?2 )
2 2
'
92c2 92c2 n
9f - 1 1 - 9§

Taking, as above, the following values for ipk

V’l(9l)=c(/1+/2)-2^-r, ^2(92)=c(/2-/1)T^-2, ^3=0

we obtain, due to eq. (24), the expression for W


W = —(fy/2 (f31 + (3 2) +

J [2/31 + 2f32q\ + 2hc2q2 (q\ - l) + 2c (/i + /2) 91 (q\ - l)]1/2 -^L=+ J [2f31 +

2(32ql - 2hc2q2 (l - q$) + 2c (/2 - /1) 92 (l - 92)] 1/2 -y==|- A change in

notation allows us to reduce this expression to eq. (13).

10.14 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)

where

f(r) = 2h + y-^. (10.15.2)


The momenta are determined in terms of the coordinates by
differentiating eq. (14.6)
Pr = r = ±y/f(r), pv = r2Cp = (3^. (10.15.3)
Function f (r) under the square root must be positive. Hence,
equation r 2 f ( r ) — 0 should have only real-valued roots 7*1,72,
otherwise r 2 / ( r ) , being negative at sufficiently small r, would
remain negative for all real r . Assuming 7*1 < 7*2 we can
write the following
T*2/ (7*) = —2h (7* — 7*1) (7*2 — r ),
where

o2
= - j ^ , rir2 “
ri+r2 * (10.15.5)
The signs of the roots are different for h > 0 whereas both
roots are positive for h < 0. We restrict our attention to the
case when h < 0. As follows from eq. (4)
7*1 < r < 7-2.
(10.15.6)
The generic names for the points of the orbit around any
attracting centre corresponding to the distances 7*1 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 Pr =Pr=f0 < 0.
Then the distance r decreases within a certain time interval (0,
£o) from the initial value 7*0 to 7*1. Thus a minus sign should
be taken in the first equation in (1) and we obtain
r0
f dr I* dr
— —. .. .. = t and to, (10.15.7)
J \/f (r)
J y/TV)
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 \Jr — T\.
Counting t from the time instant to we should take 7*1 as the
lower integration limit and the plus sign in front of the integral.
Denoting the time when the apocentron is reached by t\ we
have
r
[-*=t-to and 7-±= = 1, -
J vtf) J vm
ri ri
Beginning with the moment 11 we again have

and ) - £ = = ( , - ! „
J VTV) J VTV)
ri r\
where t<i 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 2ir. This ensures that the motion is
periodic and has a closed orbit. The time interval
1
f dr
t1-t0=h-t1 = -T= -== (10.15.8)
2
J V / (r)
ri
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
nw .ow\, xl/ ^
r = ri cos — +r2 sin — = - (ri + r2) — - \T2 ~ ^i) cos w. (10.15.9)
x

It is easy to prove that inequalities (6) are satisfied for any real
w. The substitution (9) is successfully applied in any cases
when it is known in advance that the quantity in question lies
between two limiting values (ri and 7*2 in our case). In
Kepler’s problem w is referred to as the eccentric anomaly. In the
ri +V2 = 2a, 7*2 — r*i = 2 ae

(10.15.10)
is introduced, where 0 < e < 1. Then we obtain
r = a (1 — ecos w).
10.15 Keplerian motion 579

and relationships (1) reduce to the form

rdr
/ e cos w) dw t £ Q, (10.15.12)
v^TF)

r w
dw
if : (10.15.13)
V>a( 1 - e 2 ) t -= \/l -e2 f 1 — e cos ic
j r
Vf
Here
/?„ = v V a ( l - e2), (10.15.14)
and = 0, which implies that <p 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
_ 2nay/a 2n _ y/jl
(10.15.15)
~ yfc ' ~ T ~ ayfd,'
n

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 ic = n (t - t0) = (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 ic
w ip 1 e w
2arctari W ---t a n — ,
/I + e +
ip =
V 1-e 2’
t a n — = W -----------t a n — .
2 \1—e 2 (10.15.17)
From this formula we find
e + cos ip . sm^9
cos ic = --------- , = VT (10.15.18)
sin ic 1 + e cos ip
1 + e cos ip
and the inverse formulae
cos ic — ; sinic 1 —
COS(^ sin^ : Vi" (10.15.19)
e 1 — e e cos ic
cos
The orbit is ic ’ to be an ellipse with one focal point at the
seen
attracting centre, the constants a and e being the semi-major
axis and the eccentricity
580 10. Canonical equations and Jacobi’s theorem

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 (l - e2) _ p
(10.15.20)
1 + e cos ip 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 constants 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 fixed in
the space and having their origin at the attracting centre.
Let us determine the directions of axes Oxyz by the unit
vectors ei, e2, e3. The vector ei is directed from the attracting
centre to the pericentron, e2 lies in the orbital plane
perpendicular to ei and points to the direction of (p, and e3 =
ei x e2 is orthogonal to the orbital plane. The Euler’s angles
are denoted by see Fig. 2.2. Angle £2 determines the
direction of
the straight line which is the intersection of plane O^rj and the
trajectory plane. This line is the nodal line, point N and angle
£2 are referred to as the ascending node and the longitude of
the ascending node. Angle i describes the angle between the
10.15 Keplerian motion 581

Putting now the radius vector of the moving particle in the form

a (cos ic — e) + e2 y/l — e wsin


2
r = eir cos p + e2r sin p = ei
(10.15.2
1)
and projecting it onto the axes of the system O^rjC we arrive
at expressions for the coordinates £,7y,£ of the particle which
contain the following six constants
a, e, to, O, i, a;
(10.15.22)
called the elliptic orbital elements. These expressions can
also be obtained in spherical coordinates if we use the
w=1—n e
a r’ (10.15.23)
cos w
we find, by differentiating expression (21), that the velocity
vector of the particle
v_ Gl sin w + e2 \J\ — e2 cos u?J .
(10.15.24)
The value of the velocity vector and the angle with the radius
vector r are given by the equalities
1 + e cos r•v esin w
v= cos a =- - - y/l — e2 cos2 (10.15.25)
w 1 — e rv
w
cos w ’
Now it is easy to obtain the constants a, e, wo in terms of the
initial values ro, vo, a. We have
^0 \f' 2 a + {2(5 — l)2 sin2 a;, cos wo = 2f3-l
cos
=

2(1-/?)’ e
(10.15.26
)
Here
/?=^2.
(10.15.27)
£JJL
If the attracting centre is the earth then, by replacing (i = fM in
Voo = y/2gik, (10.15.28)
582 10. Canonical equations and Jacobi’s theorem

i?o being the radius of the earth, we obtain


2
(10.15.29)

The value i>oo 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 horizontal velocity (cos a = 0) equal to ^/gRo =
0.707voo then, as follows from formulae (26), the orbit is a
circle with the radius of the earth.
Depending upon the initial velocity vq the elliptic motion
beginning on the surface of the earth (r*o = Ro) can be
divided into two types. The ” earth motions”, for which
(10.15.30)

belong to the first type. These motions take place along the
ellipses intersecting 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 = oo) corresponds to the left limiting case of inequality,
i.e. (3 = 0, e = 1, =
(10.15.31)

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 > Vqq, 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
w — ( = e sin w (10.15.32)
using trigonometric series
oo
(10.15.33)
k=1
10.15 Keplerian motion 583

is proved to be more efficient.


This representation is possible since the increment of 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

sin &C^C-

Integrating by parts and using Kepler’s equation we obtain


7T
&k — j (w — C) cos kC\o - J cos kQd (w —
7T K

—r f cos kCdw = —- [ cosk(w-esmw)


7vk J irk J v 7
dw.
o o
The functions determined by the integrals of
the type
7T
Jn (x) — — j cos (raw — x sin ic) (10.15.34)
dw;
were first introduced by Bessel for solving Kepler’s
equation. Hence,
<4(^0 • 7/-
2 00 -------sin
kQ. (10.15.35)
CLk = ~r dk (ke), w = £ 4-
2 ^^
k=l
Kepler’s equation yields
. 2 Jk (ke) .
sin w = - >e ----— sin kQ. (10.15.36)
k=i k
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
= T + ^2 bk cos
cos w
fc=l
The coefficients are given by
bo 1
? wr 1 / n
— = — cos icac = — cos w; (1 — e cos ic) dw — —-e,
2 TV J 7Tj V
00
584 10. Canonical equations and Jacobi’s theorem
7 7T
T cos ic cos kQdQJ sin sin wdw
— -—o
/ K7T
7T
7rk J cos (k£ + ic)
dw 0 .
7r 7T
1 J cos [(A; — 1) ic — A;e
nk COS [(*+1) 1C
— ke sin w] dw
sin ic] dw o /

Hence, by means of eq. (34)


1 ^ - [ Jfe-i (fee) - Jfe+i (fee)]
cos ic = — + (10.15.37)
e cos
2 /c = l
11

Perturbation theory

11.1 Method of parameter variation

Along with the given system of equations of motion


dH_ • dH^ ,\
qs = dps' Ps = - «---\-Qs (s = 1,... ,n) (11.1.1)
uQs
we consider an auxiliary (simplified)
canonical system
Qs ddH 0 . 0H
Psq (s = 1,... ,n). (11.1.2)
Ps ’ dq s

It is assumed that the Hamiltonian function H 0 of the system


of equations
(2) is chosen in such a way that we can find a general
solution containing
2n constants (11.1.3)

By analogy with Sec. 10.12 we assume that this system of


equations is solvable for crfc, fik
• • • i Qm Pi 5 • • • 5 Pm t) ■>
&k {Ql1 1 (1114)
0k =Pk(qi,--- ,Qn,Pl,--- ,Pn,t) 1,... ,n). J ^ '''
Then, as pointed out in Sec. 10.12, any of the Poisson
brackets
(ak,as), (ak,03), (0k,03)
(11.1.5)
A. I. Lurie, Analytical Mechanics © Springer-
Verlag Berlin Heidelberg 2002
586 11. Perturbation theory

is either constant (equal to zero, in particular) or expressed in


terms of the values none of the brackets depending explicitly
upon t and
variables qs,ps• By virtue of eq. (10.6.3) we can write down 2n
+ (<*„ Ho) = 0, + (0„ Ho) = 0. (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 ctk, Pk are no
longer constants but functions of time. Then equations (3) can
be deemed as the formulae transforming the new variables
OL^Pi into the old variables 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

das n das . das . da. .


dt (as,H0) + (as,H) + J2-^-Q k
dqkQk dpkPk
k=1

Here the identities (6) are taken into account and qk,Pk are
replaced by eq. (1). By analogy, we obtain (3 S and arrive at
the system of equations
n f)
as = (as,H-H0) + TQk^,
fc=3. Pks
(3s = (Ps,H-H0) + J2Qk?f -, (3 = 1,. (11.1.7)

referred to as the equations of perturbed motion. Clearly, their


right hand sides are assumed to be expressed in terms of /3 k
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
11.1 Method of parameter variation 587

Qs, Ps and u, v by as,f3s and as,H - H0. Then we have

(as,H-H0) = '£Yl -dasd(H-H0) (a , a ) +


r k
k=1 _ dar dak
r=l dasd(H- dasd(H-H0)
(<*r, Pk) + (Pk, ar)
H0) dar dpk dp,. dar
das d (H - H0) (a a ^
0/3 O/Q \PriPk)
d(3r dpk
Taking into account that
das _ das _ d^~r
~ sr’ d(3r ~
we have the equality

(as,H-H0) = J2 d{H-H0), ,td{H-H0),„ „ ;


——z------ (as,ak) +-----m-----(««>
k=1 Pk)
(11.1.8)

By analogy
n -^W..ak) + dd^WM
atH
((3s,H-H0) = J2
doth dPk
k=1
(11.1.9)

The system of differential equations of perturbed motion (7) is


now expressed in the form
d(H — Hq) { \ , d(H-Hp) ^ 0 , , da, „ '
k=1
nr — (as, a*) + ^ (as, Pk) + ^ Qfc

& = E d(H-Ho)^ d{H-Ho), , dp '


8at il3-n} + Wt i0"'M S^Q“
k=1
(s = 1,... ,n). >
(11.1.10)

The special convenience of this notation is that the Poisson


brackets are either constant or expressed in terms of a k,pk.
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
588 11. Perturbation theory

than the initial one. An approximate solution will be effective if


the auxiliary 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 quantities as,(3s, 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 a s,(3s
have the same order as the presumably small H — Ho,Q k.
Thus, as, (3S 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
t
J (a\/3;t) and t (a|/?;£), we can t replace ak,f3k in the
as = a°s+ J $s(a\0;t)dt, (3S = fi°s + J (a|0;t)dt,

(11.1.11)
o o
(obtained directly from (10)) by their initial values a®, (3® at t
= 0. 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
t t
as=a°s+J<$>s(a0\P°;t)dt, (3S = /?° + J Vs(a0\P°-,t)dt. (11.1.12)
0 0
The process can be continued, that is the values a s, (3S
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
T T
#s (a|/3) = i J *s (a|/3; t) dt, *s (a|/3) = ^ J *a H/3; t) dt,
(11.1.13) 0 0
with (3k on the right hand sides being assumed constant while
as = (a\(3), (3s='&s(a\/3) (s = l,...,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 into 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. ots,(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
(a„afc)=0, (as,pk)=6sk, (J3s,pk) = 0, (11.2.1)

which results in a drastic simplification of the form of equation


(1.10) for the perturbed motion. They take the form

d(H-H0) ,^das„
Ps ~~ (s = l,... ,n),
r; dpk
dp > (11.2.2)
d(H-Ho) , srd/3
(s = 1,... ,n).
k= 1dpk
In particular, if
Qk 0 (^ 1, . . . , 77.) ,

i.e. when equations (1.1) are canonical, then the equations for
the perturbed motion
d(H-Ho) Ps = ~ d(H-H0) (s = l,...,n)
QLa — (11.2.3)
dPs ’ das
are also canonical.
Replacing Qs, Ps in relationships (10.8.16) by as,/3s,
respectively, we can write eq. (2) as follows

d(H-H0) n 0
C TQk—,
M dPs "
-s d(H-Hp) h > (11.2.4)
das n 0
(s = 1,... ,n).
+ yQtpL
ti 0a-
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
as _d H Hq ^ ^ QkQk
_
d k=1
PS _d H Hq ^ ^ QkQk (s = l,... ,n). (11.2.5)
_
da k=1
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 -J2Qkqk,
(11.2.6)
k=1
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 consideration 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 preserve
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

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 O and is determined on the surface of the earth by the
polar angle $o (angle complementary to the latitude) and
longitude Ao- Axis Oz is directed from the earth centre O to
point O along the radius vector RQ. Axes Ox and Oy lie in the
tangent plane to the sphere of radius Rq at point O, 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
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
— ~u'
— V~JL ^ yj j i
where, as always, U denotes the angular velocity of the earth
we have
( dr' \

w • V x j = U (ii sim?o + iacostfo) • (r' x ij) = -Uy cos tf0,


and by analogy
/ dr' \ ( dr' \
u • ( r' x — J = U (xcostfo - z sin^o), u • f r' x — ) = Uy sind0.
Then by virtue of eq. (10.11.4) we obtain

T'r = - {(Pi -f- ^costfo)2 + [P2~U (xcos^o - z sin ^0)]2 +


(P3 - [/?/sirn?o)2j>.
(11.3.1)
Let us proceed to calculate the components of expression
(10.11.9)
= Ufor
2 (zsin$o — xcosfio)2 + y2 ,
the Hamiltonian function K. For m — 1 we have
(u x v0) • r' = U [(ii sin^o + 13 cos$0) x {-\2RqU sin$0)] • r'
= U2Rq sin $0 {—z sin $0 + x cos $0),
the velocity vector v0 of the pole of the start system of
asaxes equals
Vo = -hRIU sin
$0*
As it is a constant value, the term with v0 can be cancelled out
and we
obtain
- • 0 • {jj + m (cj x v0) • r' = --U2 (z sin $0 ~ x cos i90f +
2Rq (z sin i90 - x cos d0) + Rf sin2 ^0 + y2
\r£u2 sin2 d0 = ~\u2h2 + const
Z Zi
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

h2 = [(Ro + z) sin$o — xcos$o]2 + y2.


According to eq. (5.6.2) the value

II* = II — • 0 • uj + m (uj x vo) • r' = II — ^U 2h2 + const,


(11.3.2)

is the potential energy of the weight, that is the sum of the


potential energies of the gravity force and the centrifugal force.
II* is given by eq. (5.6.9) and is expressed in terms of the
distance r from the centre of the earth and angle d which is
complementary to the latitude of the actual position of the
Y
r = r0 + r' = i3Rq + hx + i2y + i32, cost? = - • (ii sint?0 + h cos$0)

and thus

r2 = x2 + y2 + (Rq + z)2 , cos t?o = - [x sin $0 + (Ro + z) cos t?o].

Here, in accordance with eq. (5.6.11)


2
R*=R0- -aR0P2 (cos^q) ,
Ro and Rq denoting the mean radius of the earth and the
radius of the normal spheroid at point O, respectively. Then

r
1+
% + xI±r-l 1/2
Ro

Az2 4z 1(x2 + y2 + z2) 16 taPa , z 1 8 z3


~pT + -----53-----------T 2 (costfo) +
16 + ‘ "
p-
u0 2 -tt
2 0 6
^ x + y z(x + y 2R ) 3y ~ 3aF>2 (cOSl?0^ +
2 2

R<y 2 R2
xyz
Here —, —, — are assumed to be small values of first order,
whereas Ro Ro Ro
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

Ro z x2 + y2 z (x2 + y2) , 2 „ , 0 , ,
2R> +3 ^(cosOoH
z2 +z(x2 +y2) 4Z " il
fig 2R$ -^-^(cosOc) R3o +' ’ ’
3 Ho
z x2 + y2 — 2z2 3z (x2 + y ) — 2z3
2

~R^ 2fl§ +
W* +
|aP2 (costfo) ^1 - + ...

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
m r\2 1 2 z 2 U2Rq 2 U2
~
3 Ro) ~ 3mRo “3 fMZ~ 3 goZ

((“"tKtJ
m\ (Ro+ \ m ( r
t
3 x2

P2 {cos do)

= aP2 (cosd) + ( ^rn - 3a) -^-P2 (cos#0),

where P2 (cos should be replaced by the following expression

P2 (cos #) = P2 (cos i?o) + ((CQS A (COs $ _ cog $ )


\ a cos v J0
x
= P2 (cos^o) + 3— cos^o sin^o-
Pq

Inserting this into eq. (5.6.9) yields

3z (x2 + y2) - 2z3


n* — goz + g$x + y — 2 z
2 2 2

go 2Pg
2Pg
2 2 U2
-/ 3zP2 (COS^O) ~ 2axcos^o sin^0 H------------ + ... (11.3.3)
3 3 go

where (3 denotes Clairaut’s constant determined by eq. (5.6.19).


dK(°) „ . dKW „ • dK{0) D 1
I=
9P, =P" V
~ dP2 ~ 2’ m =Fs’
• dK<°> . dK<® • dK^ |

£
II
1
O
C

II
0
P
'=- dx =»’ Ps=
dz = 90■ )
(11.3.5)

These are the equations of motion of the particle without the


influence of
the rotating earth. Their solutions are given by

Pi = Pi, P2 — /?2> P3 — — 9ot + /?3> (11.3.6)


2
x = (31t + a 1, y = /32t + a2, z3 = (33t - -g0t + a3.

where ol{ and (3i 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 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= U [(P1 c o s t f o - P3sind0) (P2t + <*2) ~ P2 (Pit + <*i)


cos$0+
got ((32t 4- a2) sind0 + P2 ( P3t + a31- ^got2 ) sin$
+0
9o (Pit 4- ai) + (P2t 4- Oi ) — 2 ( (3st -\- a — —got2 1
z

2R0 2 3

(11.3.7)

We obtain the following system of three canonical equations of


perturbed motion for quantities

dK(1) rr „ got , ,
OLl 00 = Ua2 cos $0 n+ (Pit + Q!
l );
op l tto
dK(!)
Oi2 d02 = —Ua 1 cos $0 4- Ua3 sin $o4- (11.3.8)
l[/c/0i2 sin^o + (/32f +
dK <*2),
Ois ~dp z rto
rr • q 200f ,. got2
= 7 = -Ua 2 sint?o - I /33i + a3---------—

0i =
and three equations for /3i
03 =
dK™ Ufacostio-^rifat
Kq + aO,
dai
dKW
da2 —Ufli cos'do + UP3 sin^0 —
(11.3.9)
It (021 + «2) - Ug0t sin
t? , itO 0
dKW
da3 -U02 sin
^o + (/^ + 0:3 - •

Replacing and /^on the right hand sides of these equations


by their initial values a® and (3® due to eq. (1.12), we arrive at
approximate solutions of the form

ax = a? + Ua°2tcos+ >

0i = 0i + Up°2tcos#o-j2-(^po1t2+ao1t)
596 11. Perturbation theory

and so on. Inserting these into eq. (6) we obtain the


expressions for the coordinates

X = a? + Pit + U {a%t + Pit2) costfo - |r (^i^r + ai^ >


y = al+Pit-U {alt + Pit2) cos tf0+
U {a^t + pit2) sintfo - -fr ( Pl~r +a°V ) - ^Ug0t3 sintfo,
Ro
<^3 + /?31 — U (a^t + (3%12)
sin^o+
2
9o 0* dot

B^Vi3~6+a3^~^4
(11.3.10)
and momenta

Pi = Pi + UpltCMSlh -+ a°lt) ,
P2 = /?2 — t//3?tcos^o + t/^tsin^o—
^(/?2^+«^)-^ffoi2sini?o, (11.3.11)

P3 = Pi- Up°2t sin ^ (/^ + <*3* - -


fib*-

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 + Uy cos$0 « Pi + U (a§ + /3°*) costf0, '
y = P2 — U (rccosi?o 2 sini?o) ^ P2 — U (aj + Pit) cosi?

o+
U (a® + Pit) sini)() — ^Ugot2 sind0,
z = P3 - Uy » P3 - U (a§ + /9§t) sind0.
(11.3.12)
The same relationships can be obtained by differentiating
equalities (10). The constants al,a2,al are the initial values of
the coordinates whereas Pl,f^,Pl are determined in terms of
the initial values of the velocities by means of the following
relationships
x0 = Pl + Ualcostio, y = Pi - Ualcostfo + Ual, z0 = Pi - Ualsintf0.
For example, let us consider a particle falling near the
surface of the earth from height h without any initial velocity,
then
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 = —U2ht2 sin^o cos$o ~ 0, y = —gt3
sin^cb
o (11.3.13)
z = h — \gt2 + U2ht2 sin2 $o ~ h — -gt2.
Z Z
Keeping the terms proportional to U 2 in the latter equality
would make no sense under the adopted accuracy of the
calculation. As a result we obtain an expression for the
eastern deviation of the falling particle. The initial conditions
for the particle thrown vertically upward are as follows
<*i=0, Oin 1, a° = 0, /3? = 0, #j = 0, (3% *0,
and, from eq. (10), it follows that
x = 0, y — Uzot2 sin$o — ^Ugt3 sin^o, z — zot — \gt2.
(11.3.14)
o Z
At the time instant when the maximum height is achieved, i.e.
when z = 0, the value of the western deviation is
y* =
(11.3.15)z
3 g
Let us consider another case of firing a particle from the
surfacea?
of =the 0,
earth
£*2with
= 0,the
a®initial
= 0, velocity v0 comprising angle
= vo cos6 cos c, fi% = vo cos 6 sin e, /3g = Vq sin 8.
Formulae (10) yield
x — vqt cos 8 cos e + Uvot2 cos 8 sin £
COS $0, '
y — vot cos 8 sin £ — Uvot cos 8 cos £
2
(11.3.16)
cos $o+
UVot2 sin 8 sin do — 7;Ugot3 >
sin do, 3
i
Placing axis O£ in the plane of throwing and axis Orj
perpendicular to it, so that the angle between axis Or) and the
western direction is £, we obtain
£ = votcos<5 + Uvot2 sin<5sin$o sine — ^Ugot3 sin sine,
o
T] = — Uvot (cos <5 cos $o — sin <5 sin $o cose) — -Ugot 3
2

sindo cose.
The latter equality determines the deviationo 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 = ^(p2x+p2y + 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 = ~^f(v), Fy = ~f(v), v= y/x2 + y2,

where x = px,y = py. Neglecting the resisting force we obtain

The equations for the disturbed motion (2.4) are thus as follows

& = —/(«), K = [ (11-4.2)

Provided that the resistance is small, we can assume


oq,... ,/?2 on the right hand sides of these equations are
constant. Then

0 0

0 0
where
v* = v (r).
For the sake of integration we introduce the new variable A
11.5 Influence of small perturbations on oscillations about the equilibrium
599
Then
v* = \j pf + (/?!] gr)2 = Pi cosh A, dr Pi
cosh XdX

and the solution of the equations in (1) can be cast in


the form
oO.^ J2 (sinh
M
x — a i+ Pit + A — sinh fi) f {p\ cosh A)
dX,
v0
y = 0%+Plt- 1 / 3° \ 2
— -gt2 + ( — \ J (sinh A — sinh^u) f (p\ cosh A)
(11.4.3)
where
P° 1
sinh no = sinh n = (/?£ - gt).
Pi Pi
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 generalised coordinates
measured fromn n the equilibrium
^ n n position by q s we represent the
T
=IEE tiskQsQk T ^ (<?i, ,qn)qsqk,
•• > (11.5.1)
s=l k=l s=1 k=l
n n
n
4EE CskQsQk T n* (^l, • • • , qn) 5
S=1 k=l
where
&sk (Qi5 ***> Qn) — Ask (</i5 •••5 qn) Ask (0? •••? 0) — Ask (^i,..., qn) &sk
600 11. Perturbation theory

ask denoting the constant component of A sk. 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)

(£).-
The power series for II* begins with terms of order not lower
than the third. It is known that the following two quadratic
forms
^n n ^n n

Q'skQsQki 2 EE CskQsQk 5
k—1
s=l s=l k=1
one of them being positive definite, can be simultaneously
transformed to forms with no products of different variables
with the help of a linear transformation, [56]. These new
coordinates are called principal coordinates. Let us assume
that the transformation to principal coordinates has been car-
ried out and ^ n use the^ same
n n notation q s for the principal
T
= 2£a^2+2££ ask (#1> • * * 5 Qn) QsQk 5
S=1 S=1 1
1n (11.5.2)
n = — ^ ^ sQs “i n*.
c -

s= 1
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 \q s\ and all the
coefficients are also positive. Of course, the coefficients a s are
positive, too. Their ratios
A* = —, (s = 1,... , n)
(11.5.3)
ds
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
dT
u
± . V--V *. (11.5.4)
Ps — =aaqa + 2^aakqk.
k=1
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

»> = 5 E(|*+<**)■ I'1-5-5)


Then the difference H — Ho is determined by the equality
1_ _ /
n
2 \ inn
H
- #0 = 2 Em - h ) + 2 a A ■ ■ ■ ’«") 9sQk + n.,
Z
s= 1 ' s/ Z S=1 k=1
(11.5.
6)
in which 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
Ps
qs = —, ps = -csqs (s=l,...,n) (11.5.7)
ds
Qs — -
COS Agt H A. sin Ast, ps
with Cauchy’s integral Ci in the form
a
asasXs sin Ast + (3S cos
Ast,
(11.5.8)
so that as and /3S are the initial values of the generalised
d(H - Ho) sin Ast
as =-----—--------Qs
(q\p;t) dPs dsXs (11.5.9)
d(H-Hp)
Ps = - das + Qs (q\p;t) cos
A3t,
where H — Ho and Qs are expressed in terms of as, (3s by
means of eq. (8).
This method of integration of this system of equations
based upon replacing the unknown variables , f3 k 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 interval,
however the qualitative character of the motion remains
obscure. The method of averaging of the right hand sides of
equations (9) is more suitable 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.
h h
—Pi + —P2
ai . a
2
h 52
—Pi + —P2
ai d 2
sin Ait fbi h
0i = - \ai—Pi +
aiAi 5
a2
II toXiy

sin A2t (h , b2
—Pi +
a 2A2 5
V°i a2

(11.5.12)
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 Ait cos A2t by the circular functions of the sum and
difference of the arguments, we arrive at the equations
ai = --—ai + -—ai
2a\ 2ai
zai
&2 [cos (Ai + A2) t - cos (Ai - A2) t] +
ZdiAi
2q2a ^1 S^n 2Ait + 2a a \ ^2) ^ S*n _
^2)
:
Pi = -7T-01 - -p-01 cos2Aif + i&iArOi sin2Aif+
Za\ Z
~^Ta2 [sin (Ai + A2) t - sin (Ai - A2)
t] - /
(11.5.13
)
——(32 [cos (Ai + A2) t + cos (Ai —
By analogy we obtain equations for a2,02.
11.5 Influence of small perturbations on oscillations about the equilibrium
603
If we replace as, (3S on the right hand sides of these
equations by constant values the result of the integrations
contains the terms
-—a°t 2a/1 ’
2a/1*’
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
ii = A*/i (xi,x2;t), x2 = m/2 (xi,x2\t), (11.5.14)

where ji is a small parameter and functions fi can be


represented
fi (Xi,x2;f) = as
(p°ifollows
(xi,x2) +
y (xi,x2)cos\vt + (xi,x2)sin\„t (* = 1,2). (11.5.15)

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, Ai d= A2. Thus, functions fi contain
components ip® (xi,x2) which change slowly in time (as their
arguments x\, X2 possess this property) and oscillating terms
with slowly varying coefficients This prompts one to
replace the
right hand sides of eq.T(15) by their mean values over a rather
lim 1 [ fi (x!,x2;t)dt = (p° (x1,x2) (11.5.16)
T-+00
1J0
and with this approximation, the equations in (15) are
replaced by the following equations
x\ = jjupi (xi,x2), X2 = IMP2 (xi,X2) •
For system (13) these differential equations take the form

13 =
‘ -?• (,=i’2)-
Their solution is given by
604 11. Perturbation theory

and substitution into eq. (8) yields

/ o
qs—e 2a« [ Q;0 COS Ast + sin \st ] ,
V a
sK ) ► (11.5.18)
_bst_
ps = e 2ds (—a®as\s sin Ast + (3°s cos Ast).

This is still a very rough approximation. It is sufficient to point


out that the term 2hq\q2 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. Returning to system (14) we assume
^ (£l;£2)sin Avt 4V) (€1,^) cos A * = 1,2,
Xu
(11.5.1
9)
where are solutions of the system of equations (17)

€i=W>i (^1,^2) -

This means that the solution is the sum of the zero-order


approximation ^ and the integral over time of the oscillating
terms on the right hand sides of system (14). During the
estimation of this integral x\,x2 are replaced by fi,f2 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 fji 2. 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).
bit
e 2ai (a® + 7-7-0^ sin2Ai£ - - cos2
^i^ 1 +
ai =
aiAi 4a? A?
b 2t
sin (Ai + A2) t sin (Ai - A2)
0 ~2ao -—— t Ai + A2 AI — A 2
a2e za2b2t
h P°2e~ cos (Ai + A2) t ^ cos (Ai - A2) t
2aia2Ai 2a
2 Ai + A2 AI — A 2
11.6 Influence of misbalance on the motion of a heavy top 605
bit
2ai Pi-----T-fii sin 2Ait — -pCK? cos
2Ait
cii Ai 4
b 2t
h\2 2a COS(AI + A2)^ COS(AI — X2)t AI +
-a2e 2
A2 Ai — A2
b 2t
h_ 2a2 sin (Ai + A2) t sin (Ai — A2)
2 t Ai + A2 Ai — A2
a2
Inserting these expressions into eq. (8) we obtain after simplification

bit e
n /?? . , \ h sin Ai7
<7i 2ai aV cos Ait H — sin Ai7 1 + -
——
>0 b2t
Pi cos Xit + h\ 2 ^a2
aiAi ai (A?-A*) (a2sinA2^ a2\2 cos X2t
(11.5.20)
The expression for q2 is obtained under the corresponding
replacement of the indices
bit b2t
hXi Pi cos Ait) + e 2a2
<72 a (A222-A?) 201
(a?sin
2 aiAi x
b2 a2 sin X2t 02 a2X2 cos X2t
a2 cos A2t H—sin \2t ) + ^a X
2 2

(11.5.21)
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 being 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 qi, q 2 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 A and
g (X = a and g = (3) as well as the angle of spin p. The
coordinates of the attached mass in the coordinate system
Oxyz fixed in the body are denoted by xq = ecose, yo = esine,
zq. The kinetic(iUyZQ
energy of the misbalance is^zVo) T (j^z^O
Ti T {^xlfO UJyXo)
[{u?x + Wy) Zq - 2u>zz0 (x0wx + y0LOy) + ...],
the values proportional to e2 are not included. Recalling the
expressions for the projections of the angular velocity
uox = — Acos^? +/icos Asin^?,
ujy = Asin^? +/Icos
Acos^?, uz = </? +
/isinA,
we find that the kinetic energy of the system consisting of the
heavy top and the misbalance is equal to
T = ^ |T4 (A2 + g2 cos2 A^ + C (p + g sin A)2 —
2ezom (p + A
sm
A) |^—A cos (p + e) + g cos A 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
n = Q\zck. • i3 + mgk • (ii#o + hyo + hzo)
= (QiZc + mgz o) cos A cos g + mge (k • ii cose + k • 12
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,
we have
II = Qz* cos A cos g — mge [sin g cos (p + e) + sin A cos g
sin (p + e)].
(11.6.
11.6 Influence of misbalance on the motion of a heavy top 607

Using eq. (1) we obtain


px = AX + mezo sin A) cos (</? + £),
— Pp sin A + cos2
A ~ mez0 (<p + fi sin A) cos A sin (</? +
e),
p^ = C (<p + fi sin A) — mezo — A cos (</? + £) + ji cos
A sin (cp A s)
Therefore, within the adopted accuracy, we have
• px mez0 / .x
A = -j - —ficVv cos (<p + e),
. Pa - Pp Sin A •/.\
V = , 2 x ~ + -77;----sm (<P + >
A cos^ A AC cos A (11.6.3)
= —//sin A + — +
mez —px cos {ip + e)-\ / x Pa ~ Pi£> sin A . , x
---------- ------sin (ip + e)
o cos A
AC
The Hamiltonian function takes the
form
H p\ , (Pn
—7- T ^^ 0 A—■—h \2 Jl+ 2Qz(j cos A cos (i
=2 -PySinXY
mezA PxPp cos A cos
/ x2 A Pa ~ Pp sin A . , N
(<p + e)----^-------p<p sm (p
o + e)
AC
mge [sin fi cos (p + e) + sin A cos g
sin (p + e)] = Hq + Hl (11.6.4)

For a nearly vertical top


P\ , (Pu,-Py>^) . #
H
°=2
A+ + -q - Qzc (^2 + a*2)
me
H-- zo \p\Pv COS (ip + e) - - Apv) pv sin (y> +
AC e)] -
mge [/xcos (<p + e) + A sin (ip + e ) ] . ^
(11.6.5)
The canonical equations of the unperturbed motion of the top

A = ^, P\ =p<pPn - Pv + Qzq\,
A
A
A = -J Ov - Pvx)> Pfi =

V=
C-—A — X’ QzcVi

Pcp= 0
608 11. Perturbation theory

reduce to the form


A + i\ = — (pM - pvA + ip\),
^ •P
(PM - PVA + ip\Y = (P/x - Pv + *PA) + <5^c (M + *A), >
x
r, ‘ ^ V \ • ftf (11.6.6)
Pcp=Pv, <fi= ~

Denoting the roots of the quadratic equation


.Q c
z
Q2 A1 A =0 (11.6.7)

by qi, #2 and the initial values of the coordinates and


momenta by , a\, (3^ and (3X we come to the equalities

jjj + iX — + {qieiq2t - q2eiqit) +


ioi \
P<p(*\
,, Qi +-Q2
\ iP\(e
iq2t
- eiqit) ,
A (qi — q2) V
.'
T (PM - 2V A + *PA) = -----iq\q2 (11.6.8)
(< ,iq2t Piqit) +
A Qi q2
n ~ PyQLx + (<qieiqit q2eiq2t) .
ipx
A(qi ~q2)
It follows from the last approximated equality in (6) that
<P = -^ + av, (11.6.9)

where a^ denotes the initial value of cp. Without loss of


generality we can assume 6 = 0 and then the perturbing
function is as follows
Hi = [(PM - Pv + *PA) e - (pM - pvA - ip\) e ] -
A lip iv

[(/i + i\) e~ lcp + (p- iX) et<p] = H x (aM, aA, a v , (3^, (3 X , *) ■


(11.6.10)
The variables ,... , /? 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
Qz P<P
(11.6.11)
Qi £ q2
~ T‘
Pu
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
P + iX = (aM + ia\) eiqit + — - P^ax + iPx) (eiqit ~ eiq2t),
Pn - p<pA + ipx = iAqi (aM + iax) (eiqit - eiq2t) +
(Pp-Pva\+iP\) eiq2t-
Then Hi can be represented as follows (11.6.12)
rnez
0 : o r.-_ /_. . \ taiq\t _ piqit\ —
i<p

Hi = 2AC iPv [% K + i«A) (eiqit ~ eiq2t) e~^+


iqi (a„ - iax) (e-<9lt - e~iq2t) e* + -!(/?„ “ + iPx) e<(®*"v) -
1 (aM + ; a A ) e ^ - ^
-m^e +
A^
(<*„ - toi) i (/8„ - /3„<»A + iftO (e*>‘‘ - e‘"‘) e-1*- -

^ W. - /V» - iPx) - e"»‘) *92^ (11.6.13)

Proceeding to construction of the canonical equations of


perturbed motion
. _dHi . _dHi k _ 3Hi • _ 9Hi
dp„ ’ aA dpx ’ ^ ’ ^A dax ’
we notice that Hi depends on (3 , (3x,a^, aA through the values
ti= fin-P<p<xx + 0x, (2 = an + ia*

and their complex conjugates CnC2- Hence,


<9#i _ dffi + dfli afli _ .fdH_i _ dffA
a/?M ^Ci aci ’ a/?A v ^Ci aci / ’
ajfi dHi dHi dHx _ ajfA ./ajrx a^A
aa„ ac2 ac2 ’ dax p* \ acx ac J V ^C2 a<2 j
0<p . 0* .c
qi = -£ = <Po, -c~q 2 = ^0 l~A

Introducing the small mez we obtain


parameter >c — o~A
an + ia A = -ixelot*(pQ 1 +A 9
Czotf
L(
P01 A g .eiVoi
C zoVo
0a ~ 0a>OLX + 10X = Xeia*C<pl 1 +
c ZQipl
(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 m 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 mG is given by formula (5.5.13). When mo-
ment mG 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 K G, the angle
between them being constant. The problem of perturbation of
the regular precession caused by the presence of moment m 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 anticipation. 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 problem 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
x + 2studied.
II = —fi v + 2h -A>( ' y - ^
iC 2

where fi = fM, f and M denoting the universal gravitational


constant and the mass of the earth, respectively. Replacing x 2
+ y2 by r2 — z1 and cancelling out in II the terms which are
independent of the orientation of the axes bound to the
satellite, we obtain
n= (11.7.1)
612 11. Perturbation theory

Let er and i3 denote the unit vector of the direction M($ 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 0, we have
z2

— = (er ■ i3)2 = cos2 0,


and the expression for II takes the form

U = ^(C-A)^. (11.7.2)

Let us introduce into consideration the orthogonal trihedron


of the unit base vectors ei, e2, e3 with origin at the centre of
the earth which is a focal point of the elliptic orbit of the
satellite. Vector ei is directed to the orbital perigee, e 2 lies in
the orbital plane and is directed to the side of motion from the
perigee to the apogee and e3 = ei x e2 is perpendicular to the
orbital plane. The perturbation of the orbital elements is
neglected, then vectors ei,e2,e3 remain fixed in the space.
Notice also that
er = ei cos a + e 2 sin a,
(11.7.3)
ei • i3 = sin $ sin?/’, e2 • i3 = — sin $ cos?/;, e3 • i3 = cos$
(11.7.4) and moreover, due to eq. (3)
cos ?? = er • i3 = sin ?? (sin ?/;cos
cosip asin a) = sin ?? sin (xj; — a).
(11.7.5)
Denoting now the major semi-axis of the orbit by a and
recalling eq. (10.1.5.15) we obtain

II = ^n2^- (C ~ A) sin2 •& sin2 (?/; — a),

(11.7.6)
where n denotes the mean motion of the satellite.
The expression for the kinetic energy of the rotational
T ~ -A (J{ +^2) + 2^3

A + 'll)2 sin2 ??^ + C (p2 + ?/>2 cos^ 1 . (11.7.7)


11.7 Rotation of an Earth satellite about its centre of inertia 613

It can be cast as follows


T A + (C-A)^ 21
= \j[A + (C-A)-?\ (7=^),
=r2

then, noticing that C — A < 0 for the body of revolution extended


along axis Gz, we obtain

M 3n2 a3 -5 3n2z a6 3
T _ s - siir sin {ip - a) < —TT C \ n uj r
1-1-
(11.7.8)
The ratio (n/tu)2 is small for satellites which enables the
application of perturbation theory, the first approximation (the
auxiliary problem) corresponding to the regular precession in
the case II = 0. By virtue of eq.
(5.6.34) , we have for the earth

= 0-0033.
o
For the solar attraction
n 1
u 366
and for the lunar attraction
n^ 1
UJ

28
Thus, the ratio |II| /T proves to be very small. For the satellite
the relationships turn out to be less favourable from the
perspective of applying perturbation theory.
It follows from expressions (6) and (7) that ip is a cyclic
coordinate, thus the corresponding generalised momentum
P<p =
C (<p + ^cosi?^ = 13^
(11.7.9)
retains constant value in both initial and auxiliary problems.
By constructing the expression for the momenta
we find
P0 Pif, Pip cos $
^= (11.7.11)
A’ A sin2 P
614 11. Perturbation theory

The Hamiltonian function takes the form


2 “1 n2
(pj, - cos A *) + ^ —5 (C — A) sin2 ft sin2 (ip — cr),
sin2 $
(11.7.12)
where the constant term /32/2C is omitted. The Hamiltonian
function for the auxiliary problem equals
(p^-Py cos'd)
2
A
(11.7.13)
A sin if
2

which corresponds to the regular precession, as pointed out


above. The coordinate ip is cyclic, hence

^ = -^=°, P* = /V
(11.7.14)
The vector of the angular momentum of the body can be
written in the form
K = A (n'd + nfip sin + C13 (jp + ip cos ^ ,
where n, n',i3 denote, as usual, the ”half-moving” trihedron of
the unit vectors of the nodal line, the perpendicular to it in the
ujn = $, LJn' = ip sin $, (J3 = Cp + ip cos if
denote the projections of the angular velocity vector on these
directions. By virtue of eqs. (9), (11) and (14) we obtain
V~ Py cos if
K = np# + n' + (11.7.15)
sin if
Since the moment of the external forces about the centre of
inertia is zero, i.e. mG = 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
(/s, -/>*<*»») <1L7-16>
Let us recall that the vector of angular velocity of the ” half-
moving” trihedron of unit vectors n, n', i3 is given by

u0 n if + riip sin if + 13^ cosif (11.7.17)


COS if
-1 (nptf + n - Pleas'd . <P COS'#

sin if '-----------------
sin21?
11.7 Rotation of an Earth satellite about its centre of inertia 615

and thus
P$ - Ay COS'# 3$
n= ct>o x n n ~V'
A sin d \ sin d
cos'd,
— - n--------- .Py22 ^
CQS
n -w0xn = i3r_._P# (11.7.18)
A A sin d
di3 . ,p-0 , P^-13 cosd
-=u,0xi3 = -n- + n———.
We find
i> , • , (p* -0 COB#Y , , Pj-PyCOB#
V
di3
K =p#n + p# n + -z-—-------------------- n + -----n
+ /?, 0,
* dt
V sini; / sin d (11.7.1
9)
which can be easily proved by inserting expressions (11) and
(16) for the derivatives d, p# and the unit vectors. As follows
from eq. (15)
i3-K = /J„ e3-K=^.
(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 e 3 along K and denote
0 = 0O, fa = K, (3<p=K cos 0O, (11.7.21)
and, by means of eqs. (11) and (9), that
Py - P& cos 0Q A P< — 4/ COS 00
A-C
$ $> AC K cos 00-
(
A sin 0o
2
A ’
C (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 0 of
axis i3 at time instant t is
Kt
9 = 2k tan — = 2k tan ——, (11.7.23)
2 2A
provided that the initial position is at t = 0. By virtue of
Rodrigues’s formula (3.1.11) we obtain

is = i3 + 2 tan ^k x ij] + 2 tan2 ^k x (k x ig) (11.7.24)


\I>
1 + tan2 — L
616 11. Perturbation theory

Taking into account that K is a fixed vector we obtain, by


means of eq.
(15),
k
=^ (3^ - py COS^Q
n°/3tf n
/0
i§A (11.7.25)
sin $o 3A"V
where ^8 denotes the initial value (at t = 0) of momentum and
the index zero denotes the initial values of the other quantities.
Carrying out calculations we have
i3 = cos ^ + k x i3 sin + (1 — cos kk • i3

or
costf0 sin ^ P^P# /i lTf\'
fit - P<p
13 = n —----------------+ 1_cos^) +
sin VQ K K
cos^q)
sin 9 Py (P^ ~ Py
n P-&

+ (1 — cos^)
K sin $o
2
K
P2
cos ^ ^ (1 — (11.7.26)
cos \P) z
K
By means of this relationship it is easy to obtain the
expressions for Euler’s angles d and xp 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.
i3 • eiRecalling
= sin d sinthe formulae
is • e2 = — sin d cos i3 * e3 = cos $,
n° • ei = cosV’o? n' • ei = -sin^ocos$o;n° • e2 = sintp0?
n'° • e2 = cos 'ipQ cos $, n° • e3 = 0, n'
• e 3 = sin
we arrive, after elementary transformations, to the following expressions
PipP P# sin P'ljjPi
cos'd = cos ^cos $o tj, sin^ + p (11.7.27)
K
2 K ~K?
.o• / .o. / T , sinv[r (P^ -P costio , ,
sin i;sin^ = sin VQ sin 'ip0 cos W H-------— ( ------ cos
^0+ K sin
^o. / X i-cosVE' (Q . Pv-Pih cos
cos^o . \
P# cos Sin ip0) H----^2- - -P<p (P# ^0 H---------------Sm ) ’
(11.7.28)

sin d cos ip = sin $o cos 'ipQ


cos ^ +
sin ipQ + p# cos $o cos V’o J +
sin
K ^ ( P^ ~ Py cos #q
1 — cossina , $o
a- , , Pv-
PrpCOS# 0 cos^o ) •
^ (3 -0# sin ip 0 + -------------
sin $o
K~2
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 ip. Of course, this
derivation, based on the simple formulae for the regular
precession, is essentially more simple. Let us recall also that
A’ a2 , {Pi,- PpCOstio
Kz + 0 <Pm (11.7.30)

In addition to formulae (27)-(29) we write down the relationships


K cos sin ^ ^ sin PQ cos
P*p=(3^, P0 sind
(11.7.31)
the second one being obtained by differentiating expression
(27).
Thus, we obtained the generalised coordinates and
momenta for the regular 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 /?$, $o> V’o
^o = dU V’o dU dn • _ _dn <9?V
(11.7.32)
d/V

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 /3^ in eq. (15) by p^ (this
momentum is no longer constant) and using formula (12) we
have

2
n3
K = 2AH + /3 — 3n —A (C — A) sin2 $sin2 (ip — a).
2 2

(11.7.33)
In as much as the energy integral exists we can write
d a3
2KK = 2K K = -3n A (C - A) —sin i9 sin
2 2
(ip - a).
2
(11.7.34)
at r6
The canonical equations of the initial problem can be written in
the form
gU a3 '
= (Ptf)o - -Qfl = (Mo - 3n2^g (C - A) sin?? cos•dsin (ip - a),
2

(11.7.35)
618 11. Perturbation theory

where (p$)0 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)

. a3
K = —3n2 — (C — A) sin $ sin (ip — a) [n cos # sin (ip — a) -b n' cos (ip
— a)},
r

since the result of the other differentiations is identically equal


to zero due
to eq. (19). We obtain
. a3
K • ei = —3n2 — (C — A) cos # sin # sin (ip — a)
sin cr, ► (11.7.37)
. a2 3
K • e2 = —3n — (C — A) cos # sin $ sin (ip — a)
cos cr,
. a3
K • e3 = —3n2— {C — A) sin2 # sin (ip — a) cos (ip —
a).

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
2-7T 2-7T

K-e, = 5^3„2(C A) i sin a da J sin •& cos # sin (ip — a) .


o o
(11.7.38)

In order to gain a more precise picture of the perturbed motion


it is necessary to take into account the long-periodic
oscillations restricting the analysis by averaging only over
variable and neglecting the change in the quantities
depending on a since they do not change considerably over
the period of variable \k.
As follows from relationship (34) the value of vector K is
constant even after this averaging. Averaging the right hand
11.7 Rotation of an Earth satellite about its centre of inertia 619

obtain

2 KK =-
o
X
o 3
n
- - -A (C — A) — sin2 $ sin2 — cr) — =0,
Q^2

- - -(11.7.39)
27T A r xT/_n
since sin <ip sin d and cos ^ sin $ are periodic functions of 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 a3/r3 by
-g (1 -f ecoscr)3 . (11.7.40)
(1 — e2)
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

(11.7.41)

o
and

sin 2 — a) da = 0. (11.7.42)
o o
The projection of vector K on the normal e 3 to the plane of
the circular 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 #0,/?$, should be taken
as
620 11. Perturbation theory

where the terms yielding zero after integration over 4/ are


omitted. Having carried out the integration and further
simplification, we have
T> 3n2 / 3/3 \ /f^-f^cosOo cos ip0 - -j^r sin ip0
K.ei = _(C_A)M__

(11.7.43)

and by analogy

K • e2 = * (C - A) (l - Z£\ % *>*, + ff «*,)

These relationships can be easily cast in a more transparent


form. By using eq. (25) we find

. P&/
rv , Pip-P* c°stfo .
K cos^o + K sin $o sin^o,
k- ei = —
(11.7.44)
, P* . ; Pip-Pi, COSI?0
k • e2 = -ante - »S</V

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 ©o
and Xo> respectively. Then we obtain
3n2 Q _
K ei =------ - - -—— cos©o (l — 3 cos2 ©o) COSXQK •
e 2, 4 CV0 ► (11.7.45)
cos ©o (l — 3 cos 2
©o) cos • ei,
2
3 n C-A
K e2 = X Q K
4 CV0
where r$ 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
3^2 fi _ A
£2 = —- - -—— cos©o (l - 3 cos2 ©0) cosxoe3>
-----------(11.7.46)
4 Cro

directed parallel to the normal e 3 to the trajectory plane, [4].


Thus, the satellite executes a regular precession about constant vector K which pre-
cesses with angular velocity £1. The effect of perturbation on the regular
precession is explained by the non-sphericity of the inertia
ellipsoid of the satellite (A ^ C).
11.8 Equations of the perturbed Keplerian motion 621

1 1 . 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
F
o = -^e-> (n.8.i)
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 Fo- The problem of perturbation 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 motion. 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
difficult 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 unperturbed
motion. Expressing the eccentric anomaly w in these
a (1 - e2) M1
1 + e cos v = a \/l — e2 [ere sin ip -b (1 -b e cos ip)].
p’ (11.8.2)

Here denotes the unit vector of the perpendicular to the


radius vector r directed to the side of increasing ip and lying in
the plane of the unperturbed motion, and e% = er x e^. The
trihedron er, e^, e3 is obtained by rotation of the trihedron
ei,e2,e3 about e3 through angle ip, with ei 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 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 differentiation of r
622 11. Perturbation theory

Let UJ denote the vector of the angular velocity of the


trihedron ei, e2, e3, then the angular velocity of the trihedron er,
e^, e3 is u? + e3<£>. According to the formulae for differentiation of
unit vectors (2.7.5) we obtain
er = (u? + e (p) x er = (<j3 + <p)
3 - c^e3, '
= (u? + e (p) x = u e - (<j3 + cp) er, >
3 r 3 (11.8.3)
e3 = (UJ + e3y?) x e3 = ujp& ~ uj &p. r r J

Here ujr, , o;3 are given by formulae (2.9.3) expressing


projections of uj
in terms of the derivatives of Euler’s angles
• A • • •
ur = UJi cos + 6^2 sin ip = — cos u + i 2 sin i sin u,
atdi • - (ll 8 4j
— sinw + . A

ujp = — o;i sin + 6^2 cos pj SZsmzcos u, \


•J
——
where
u = (J + (f (11.8.5)
denotes the angle of the radius-vector of uj with the direction
on the ascending node.
In the unperturbed motion the trihedron er, e<p,e3 has angular
velocity e^(p° which is different from the component e$(p of the
vector of angular velocity of this trihedron in the perturbed
motion. Hence, in the unperturbed motion formulae (3)
become
e° = (p°e^, e° = -<p°er, e° = 0. (11.8.6)

Differentiating expression (2) for vector r with respect to


•oe a(! 2
) a (l e2)
V <P ~ +e —~^(p°e
1 + ecos (p (1 + ecos (p)sirup +
/. . 0\ d . d . d a (l — e )
2
+
^P~tp ^ + ed~e+ad^ 1 + e
a (l — e )
2
cos (p
+3 + ( p - Cp°) - u j ^ e z ] , (11.8.7)
1 + e cos
<p
the two first terms implying the derivative r*o calculated for
the unperturbed motion, and they give value v defined by
formula (2). This is easy to prove if we replace (p° by
0 = yjna (1 - e2) p (1 + ecos
^ rp2 (11.8.8)
<p) O a (1 —
2

e) ’
2 3/2
11.8 Equations of the perturbed Keplerian motion 623

determined by eqs. (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,e(p,er should vanish. This yields the first set of equations
for the perturbed motiondi . A • • ^
— (11.8.9)
— — smu + SJsmzcoszz = 0,

Cp — CfP — CJ3 = (jb° — + & cos z^ ,


(11.8.10)

cusesin p + ^2 (^e cos ^ ^


cos
V) — — (1 + e cos p) = 0,
(11.8.11)
where in eq. (11) value ip — is taken from eq. (10).
Let us proceed to differentiation of vector v. Using eqs. (6)
and (8) we construct first the expression for v° which is the
derivative of v in the unperturbed motion
V° = —+ <9ve3</?° x V
op
= <p°\/— u L ,, [erecos<p — evesmp + e^esin^ — er (1 4-
ecos<p)]a
_ -o /M 1 „r
- *Va ~ er = F0. (11.8.12)

Next, we construct the equality


V - v° = F = Frer + + F3e3.

(11.8.13)
Because of eqs. (9) and (10), the expression for u + e^ip
takes the form
u) T e%ip = uorQr T (,b e3,
0
+-a?+^+^°e3)x v - —y - ^°e3 <9v
V—V xv
da dp1
, d . d . d\ ul
— I — 3— dp
1- e——|- a—— ] v + cu e3 W — (1 + ecos ip),
de da a Vl - <
UJ r

(11.8.15)
and the second set of equations for perturbed motion becomes
di •. .. [a y/l — e2 _
ujr = — cos u + S2smz sinu = ,-----------1% (11.8.16)
at v (i 1 + e cos ip
624 11. Perturbation theory
e (cos (p 4- e) a, . ,n
,
1 _ e 2-----^ (1 + ecosp) + a;3esin^ = 1 - e2i^, (11.8.17)

e. a.
------ sin — —e sin p — uo^e cos e2Fr. (11.8.18)
<p
Equations (9) and (16)
di _ l~ayield a A/1 — 1
dt \ fjL 1 + e cos F3 cos u , O sin i = A — F3 sin u,
//1 + e cos
fa \/T — e2 (11.8.19)
and eqs. (11), (17) and (18) render
• A------2/r • , e + 2cosy? +
ecosV <p > (11.8.20)
e = . -Vl - e2 Frsin<p +---------—-------------—
a la" 1 2 a
V M v^l-— [Fre sin + (1 + e cos p) F[
<P\ > (11.8.21)

a \J 1 — e2 ( ^ 2 +ecos . \ . ^
----------( —Fr cos 99 + -------—F„, sm(y9 ) .
----------(11.8.22)
/ (i e \ 1 + e cos p ^ )
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 Cl given by the second equation (19) and formulae
(4), (5) and (8), relationship (10) F3CL
is transformed
2
sin?/cot i (lto—the
e2)2 form
p (1 + ecos 1 ' “Q • (11.8.23)
t1 (l + ecos<^)
(p)2

Using thisa (1 expression


— e 2) 2 we transform to the independent
variable u in eqs. (19)-(22) and obtain the equations for
perturbed motion in the form
di_ Fs cos u dVl Fs sin u
1 ^ ^ - C (1 + ecos(p) sini ’ (11.8.24)
JT

de (_ . e + 2 cos p + e cos2 (p _ \ ,^x


s =G(FrS,„v+-----------------1+^cosy VF«). (11.8.25)

1 da G
- e2 \Fre sin ip + (1 + e cos <p) F^ (11.8.26)
2a 1
du
11.9 Perturbed motion of the centre of inertia of the Earth satellite 625
d G 2 + e cos sin u cot
w e —Fr cos tp +cp sin cp eF3 (11.8.27)
i 1 + e
du
1 + e cos
Here for the sake of brevity cos <p

la2 (l — e2)2
(11.8.28)
r M (1 + e cos </?)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 related to the geometric constructions
requires uncomplicated calculations suggested by the method
of parameter variation.
Angle ip in eqs. (23)-(26) is determined due to eq. (5), as
the difference of angles u and UJ. If Fr, 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

1 1 . 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 noncentral 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 @3 = C and Oi = 02 = A into eq. (5.5.6) yields
the potential energy of the additional3force of attraction due to
C-A 2 ^ r3 V 3 r2
(iC — A (C2 + rf 2C2)
2 r5
The origin of the coordinate system O^rjC is taken at the centre
of the earth,
axis OQ being perpendicular to the equatorial plane.
Replacing C — A in
eq. (5.6.23) we obtain 1
3 r3 (11.9.1)
n=—
where e = 0,00164 is a constant of the earth shape. Its
smallness leads to
626 11. Perturbation theory

We have
F = - gradn = (l - 5^) er - 2e^k,

(11.9.2)

with k denoting the unit vector of axis 0(. Noticing that


r ^ k • er = sini siniq k • = sin i cos u, k • e3 — cosi,
we (11.9.3)
find
Fr = — e- 4 (l — 3 sin i sin ?x)
2 2

r
Fm = -£ » sin i sin 2u,
2
(11.9.4)
r, »Rl • R .
b3 = — e—— sm 2i sin u.

The value T is given by


Rl
T = 1 + 2e- cos 2i (1 + ecos (p) sin2 u. (11.9.5)
' (1 — e Y

It will be needed later, however we can take T = 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
di E\ .
system of equations
x _
— = —— sm 2% (1 + e cos ip) sm (11.9.6
2u, du 2 )
dft -£i cos i (1 + e cos ip)(l — cos 2u),
du (11.9.7

de
du = -£i (l — 3sin2 usin
2
i) sirup (1 + ecos cp)2 +
2
(11.9.8)
1 da £i (l — 3 sin2 u, sin2 i) e sin ip (1 + e cos <p)2 -f
2 1 — e2
a du 2 / \ ^ sin i sinecos
2u (1 +
ipy (11.9.9)

du £i (l - 3 si
sm u sm i) (1 + e cos (p) cos <£-
; 2
du eL
2
sin2 i sin 2u sin ip(2 + e cos y>) (1 + e
cos y?) + 2e sin2 u cos2 i (1 + e cos (11.9.10)
(p)] ,
11.9 Perturbed motion of the centre of inertia of the Earth satellite 627

where
Rl
£\ — £- (11.9.11)
(1 — e2)
Restricting our consideration to determination of the secular
perturbation 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, 27r) of the change of
variable u. While averaging we deem the elements a, e, z,o; of
the elliptic orbit to be constant whereas angle p should be re-
placed by u — UJ. For example, let us consider the process of
averaging the value in the square brackets in eq. (8). We
, 3.2. 3 n.2.
1- -sin i + - cos 2u sin sinp + e sin 2p + -e (sin p + sin 3p)+
2

sin2 i sin ^e + ( '-e1 + 2 ) cos p + cos 2p+^- cos 3p


2u 3 : sin2 i (cos 2u sin 2p + sin 2u cos 2p) + ..
3
—e sin2 i sin (4u — 2UJ) + ... .

The dots denote terms whose mean value is zero. A similar


calculation yields
da Q de du ’ „ di „ (11.9.12)
du °’ T = °-
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
— = —£i cos z, = “£l (l — 5cos 2 i)
(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
The change in the sixth elliptic element, which is the time of
pericentron passage, is much more elaborate. Let us notice
first that relationship
(10.15.17) for the true anomaly p, in terms of the eccentric
anomaly w, is an integral of the equations for the unperturbed
motion containing the three constants e,a,£o (the latter two
constants are due to Kepler’s equation (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
628 11. Perturbation theory

Bearing this in mind and differentiating eqs. (10.15.17) and


(10.15.17) with respect to time we obtain
= vT w e sin ip . C + e sin it;
(p + ----w = (11.9.15)
1 — e cos w 1 — e2 1 — e cos it; ’
where ( = n (t — to).
Removing w from these equations and replacing cos it; and
sin it; by eq.
(10.15.18) , we obtain
£(l + ecos<p) esin<^
if : + 1 — e2(2 + e cos ip). (11.9.16)
(1 - e2)3/2
By virtue of eq. (8.23)
• • . A tLA /m(i (1 + e cos ipY du
= IX ~ UO = 'll I 1----— ] =du \-------- T1 o
V J V a a (1 - e ) 2 3/2
du
/i (1 + ecos<^)r
2 duo
i + (r -1) (11.9.17)
« a (1 — e2)3/2 du

and substitution into eq. (16)


yields duo de sin ip (2 + e cos <p)
£— n = n (11.9.18)
du du 1 -e2
By means of eqs. (5), (10) and (8) the right hand side of this
equation is
reduced to the form
(l 6 COS (p) r/-ir»*2 o
2\ / o \
sin
77,51 i sin 2ix sin <p (2- +sme usm
- - -(1---2)— LV-*- ^ ' cosV<p)] (2e .— c o s — ecos +
On using relationship (23) and Kepler’s equation, the
expression on the left hand side is transformed as follows
C —n = h(t-to)-n^
dt (11.9.20)
1, . . dn dto (1 + ecos ip)2
— (it; — esmw) —----n ——
n du du a (1 — e2\3/2
2
) *
Noticing that
1 dn 3 da
n du 2a (11.9.21)
du
and replacing w and sin it; by means of eqs. (10.15.17) and
(10.15.18) we
come to the following
= h Up) 2 differential— eequation
arctan J1tan ^ + f2 (p)
n (11.9.22)
du
11.9 Perturbed motion of the centre of inertia of the Earth satellite 629

in which
3^1 /i , \2
h W) = — sin i + - cos 2u sin i J e sin p+
1-e ---- 2 (1 + e
cos p)
(1 + e cos p) sin2 i sin 2u\ , (11.9.23)

{(j-frf.+l
h (<p) = ~ £i : (1 + ecosy?) c o s 2u s i n 2 i x
e\/l — e2
[e (2 + e ) — cos p (l — e ) — e ( l + 2e ) cos2 p] +
2 2 2

sin i sin 2u sin p [2 + e + e c o s ( l + 2e2)]} .


2 2

(11.9.24)
The mean value of function /2 (y>) over the period of the
satellite revolution
27r 3.2A/ e2\ 3e2 . 2 .
^ J f2{v)dv £i 1
'
1 — — sin i J I 1 + — I H—— sin i cos
2a;
(11.9.25)
Function /i (y>) is represented by the trigonometric series

/i (^) = Y^t2 ^ (a/, sin kp + 6/. cos kp). (11.9.26)


fc=i
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
a\ = (l — ^ sin2 ij e ^1 + (3 + e2) sin2 i cos

2a;,
► (11.9.27)
<22 = ^1 — ^ sin e + ^1 +
2 2
sin ^ 2 cos
2a;,
/ 3 . 2 A e3 9 / es2m\ .1 c2 o.s ~
ci‘3 = ( 1 — — sin zl — + - e l l + — 1 2a ; ,
3 5
We can apply the equality

!\- e 1=^/1 dip

2 arctan ^ /---tan
1+e + e cos ’ ip
630 11. Perturbation theory

to estimation of the
integral
1 f 2ir 11 - e (f
— (ak sin k(p + bk cos kip) 2 arctan W --- tan
—dip
27r J V 1 eZ
o
27r 27T
\/l - e2 [ -———- [ (afe sin fc</? + bk cos dtp
J 1 + ecosip J
0 l\)
2ir 2ir
k cos kxpdxp ~T^f T sin k'lpd'ip
1
+ e cos xp + e cos xp

By means of substitution e1^ = cr, the


integral
ih1 27T p^d'lp
Ik + e cos xp

is reduced to the form


d 2 f o do
k
All okdo
Ik 2iri j ecr2 + 2a + < 27ri e / (cr

where integration is carried out over a unit circle. The roots of
the denominator are given by

<7! = A (l - \/l - e2) , <72 = A (l + a/1 - e2) .


|(Ji| < 1 and | <721 > 1. By Cauchy’s theorem we obtain

2 a’l
h= = (-i r
e(cri -<72) eky/l — e2
It is a real-valued number,
henceefc\/l — e2 27T
-/- cos kxpdxp
= (-i r 2tr7 1 sin kxpdxp = 0.
27T J + e cos xp + e c o s xp
1 (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
U60 _ Qfci (11.9.29)
dto 3si ^ak \, (l - \/l(-1)'
- e2) 1
Ofc k=l
72
du 1 —£i e2 k 1 — ^ sin2 A ^1 + —-'j + ^e2 sin2 i cos 2LU
11.10 Variational equations 631

where the coefficients a*, are given by eq. (27). Limiting our
analysis to terms of first order in e we find
- = —£i (1 + 3e) (l — 3 sin 2 UJ sin2 i) .
(11.9.30) 7
au

1 1 . 1 0 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 substantiates 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 equations referred to as the
variational system. Its integration is simplified by the possibility of
obtaining certain particular solutions whose number coincides
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
qs = Qs (qi,--- ,Qn) (s = i,...,n) (n.io.i)
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
qs = fs (t - t0, Ci, . . . ,Ck) (s = 1,... ,n)

(11.10.2)
which is the unperturbed motion. The initial values of
variables qs in the unperturbed motion are designated by /°,
that is
at t
= to qs=fi = f,(0,C1,...,Ck). (11.10.3)
The difference between the perturbed and unperturbed
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
Qs = Qs {Qh • • • 5 Qn) T ( ( / l ) • • • 5 Qm t) f 5 • • •
5 ^0 5 (11.10.4)
with fi being a small parameter. The perturbed solution is the
solution of this system satisfying the initial conditions
= t0 qs = fs+x°s>
at t (11.10.5)
where /° and 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

1,... ,n).
q a = f s ( t - t 0 , C i , . . . , C k ) + x s (s = (11.10.6)

Differential equations (4) are rewritten in the form


fs T X Qs (fl T 3?1, • • • 5 fn T X ) T {fl T • • • 5 fn T Xm •
s
=
n

(11.10.
7)
Functions fs are solutions of system ( 1 ) , i.e.
fs = Qs (A,... ,/n),

(11.10.8)
so that
%s = Qs {fl X\, ... , fn T xn) — Qs (/i, • • • , fn) T
H$s {fl + Xu • . • , fn + Xn) {s = 1,... , n) .

(11.10.9)
£o ^ t < to + T. (11.10.1
1)
11.10 Variational equations 633

Then we arrive at the system of equations

xr + n$s (fi, t) + Rs (s = 1 , . . . , n),


o
(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

(11.10.13)

This can be deemed as being a small value of higher order


since it is multiplied by /i. The first-order approximation is thus
the following system of inhomogeneous linear equations

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 theorem on continuous dependence of solutions of the
system of differential equations on parameters and initial
values: given a system of differential equations
Qs = Qs ( t f i , . . . ,np) ( s = 1 , . . . , n), (11.10.15)
whose right hand sides are continuous with respect to all
arguments gr, t, fi{ 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 and have continuous partial
derivatives
with respect to to,q^,jJii in a sufficiently small domain 8, see [33].

be the solution of the system of differential equations (15) for


fixed values
$ of parameters fi{ and let qs = /° at t = t0. Then, for sufficiently
for
small
fs | ^ Vs 5 | l^i | ^ Mi (s 1 , . . . , n, i 1,... , p) ,
(11.10.1
7)
634 11. Perturbation theory

the solution of the above problem

in a certain time interval (11) will satisfy the inequalities

(s = 1,... ,n). (11.10.19)


It is important to mention that values of r]s and Mi determined
through £s depend, in general, on T. The greater T is the
smaller these values are. If r]s 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 admissible to consider for
Q = 0, q = 0,

where q is measured from the vertical position. Taking, for


instance, e = 1° and i: = l ° / s e c we can determine such small
values of the initial deviation go and initial velocity qo that
within the time interval T = 1 hour inequalities (19)
\q\<£, \q\<£
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 equations and,
on the other hand, illustrates the fundamental significance of
the concept of Lyapunov stability.

1 1 . 1 1 On integration of variational equations


Let us consider the system of linear homogeneous equations
(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 Ci of solution (10.2) of the system of
equations for the unperturbed motion (10.1)
^) = IQ (* = !>••• >«;* = !.••• >*)> (11.11.2)

as well as its time-derivative


vik+1) = Is
(11.11.3)
are partial solutions of the system of linear homogeneous
equations (1).
•• A dQs ■
dCi ~ “ dfr dCi ’ Js ^ dfrJr
This can be recast in the following form

(11.11.4)

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 = fs(t- to; (11.11.5)

q°n) ,
7s = fs (0,9?,. •• ,q°), dh —3 is: (11.11.6)
t=t0
being the Kronecker delta. From relationships (2) we obtain
6is
the complete system of solutions of the homogeneous system
(1)
= 0 (a,z = l,...,n). (11.11.7)

It satisfies the initial conditions at t = 0


r]^(0) = 6iS (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

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
x
n
s = Ylxkrl(sk) (t) (s = l,...,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)
can be readily found by means of the method of parameter
variation. We
take
n
xs = Y,Ck^T)sk)(t), (11.11.12)
k=1
with Ck being determined from the following system of
equations
n
=/i$s(/i,... ,/„;f).

(11.11.13)
k=1 dn^$rArfe(*), (ii.ii.i4)
It follows from eqs. (13) and (9) J r—1
that

*(t>=^ (-/£(©
\ o S~L
where Ark (t) denotes the algebraic adjunct of the element of
the r — th row
n
x. = E ^ w
fe)
x°k + n Arfe (f) exp

(11.11.15)
11.12 Equations for perturbed motion of a 637
particle

FIGURE 11.1.

1 1 . 1 2 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)
9

Let us refer to the trajectory C of particle M under this


motion as supporting curve. Further, r, n, b denote the unit
vectors of the natural trihedron, p and a 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 motion.
The position of point M on the supporting curve is
determined by the position vector r (a). Let the normal plane P
of the supporting curve passing through point M* intersect
this curve at point Mi with the position vector r(a + e). Let
T i , r i i , b i denote the unit vectors of the natural trihedron of the
supporting curve at point Mi, whilst v and f3 denote projections
r* = r (a + e) + mi/ + bi/3. (11.12.2)
638 11. Perturbation theory

Using Frenet’s formulae (2.18.9) we obtain


1 £2
r (cr + e) = r (ct) + sr' (<r) + -e r" (a) + ... = r (a) + er + —n
2
2 p + ... ,
ni = n - -t + |;b+,... , bj = b - En + ...

and, thus within the accuracy of the squares of small values,


we have

r
* =r + TE (‘ - ?) + " (" + T„ - t) + b 0*+ y) ■ <u'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 + r ( e - y..) +VG
v
( i - + ^VG
n
+ ^ ) ,G (. EG 0g\~
- ^ ) +, ( / 3VG
b

W* = W + T £--------1 2 ~P--------( ^ H---------777 +


p p2 p p\ p T) \
.. EG EG , EG 0G 0& _ 0G
1 2

p p2 H p T T2 T
G (. V(j\ G ( • VG x
-e{£-j)-f(e + lf) +

?+ ' + Y + °T7J-T
' "(11.12.4)
where a prime denotes differentiation with respect to arc
length g. Casting
the vectorial equation of motion as follows
w* - w = F* - F,
we obtain three variational equations

S - 2— - -Fr + ^-Fn - -Fn + = F* -T Ft, (11.12.5)


ppppT ’

EG E EQ V
v + 2- - -1—F -----—Fn-- - -Fn —
T

P P P P
2^ -IFT + E^I ((3T' -v) = F* - Fn, (11.12.6)

/? + 2^ +
V
-Ft + ±Fn - ^ (isT' + P) = Fb*, (11.12.7)
11.12 Equations for perturbed motion of a particle 639

In these equations a and a2 are replaced by expressions


from the equations for the unperturbed motion (1). As follows
from the derivation, F*, F'* and F6* denote the projections of
force F* acting on point M* in the perturbed 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
q1=a + e, q2 = v, q3 = (3 (11.12.8)
can be viewed as curvilinear coordinates. The unperturbed
motion corresponds to the values = cr, q% = = 0- The basis
vectors of this
coordinate system are determined by formulae (B.4.4)
r =
dr\ ( v \ (3 v
‘ a? = ’ A 1 - f t J - n i 5 T + i7b'' r2 = ni' r 3 ’ b l -
Then, applying Frenet’s formulae and neglecting the terms of
second order we find
v + n|£-|)+bi,
ri = r ( 1 -
PJ \P
£_£
r2 = -T- +n + b— (11.12.9)
P T
r3 = -n- + b.
These formulae can be obtained by differentiating expression
(3) with respect to e, v and /?.
The covariant components of the metric tensor are
2z^ (3 v
#11 = 1- - -> 912 = 913 = ,
P T T
922 = 1, 923 - 0,
933 = T

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 2v.11 = 1 + -, gu = fL, g^ = -£
P T
„ 22 1, g23 = o,
a33 = l.
Using this result and formulae (B.1.14) we obtain vectors of
the co-basis
ri = r (i + ^ + n^,

r2 = T
(?~?)+n+b?’ ’ (11.12.10)
V£ ,
r, = -fr--„ + b.

j
640 11. Perturbation theory

The vector F* is represented in terms of the covariant


components F* by expression (B.2.4)

F* = F*rs = F* • r r .
s
s (11.12.11)

In the case of potential forces

F* • <5r* = F* • ra6q8 m
dU
=•

the potential energy II being considered as a prescribed


function of the curvilinear coordinates. We obtain
*
F 3= F * = _dn s
dqs’ dq (11.12.12)

For the unperturbed motion

an
(F*)0 = F = FTT + Fnn=-(—v3^
an\ /an Aa?VoT+
\d<i2
and thus
<n i2 i3)
dU0 ■f" - ( £ ) „ " * ■ (§)„-°- - '
da

where IIo = II (#o, <#)> tfo) = n ( c r , 0 , 0 ) implies the potential


energy on the supporting curve. Returning to expression (12)
for force we have
on (dn\ / d2n \ ( k k,
dqs \dqs) 0+\dqkdqs) 0^9 ?
o)+---

or, taking into account eq. (13),


an / dFr dFn — = -\FT + £—+
da da
dU
Fn+e^.m) „
dq2 da \ dv2 ) 0 V \ dvd/3 ) oJ
an _ ( a n \ . „ (d2n
2
dq3 v
~ +(
{dvdfi) \d(3
0
2
+
(11.12.14)
11.12 Equations for perturbed motion of a particle 641
Due to eqs. (13) and (14) we have

™(11 + £) ®(|_£
F* ____dq \ p) dq2 \T p
_ (. dFr dFn T
~ Fr ( 1 H — ) + £ -------V v—-----b Fn
\ p J da da
an e an an e
F* = —
M
n dq1 p dq2 dq3 T
= FT- + Fn+£—- 2)^- [o/jdis)
an e an _ e
FZ = ~ ~dcpT ~dq*=T
(11.12.15)

These expressions should be inserted into the right hand


sides of eqs. (5)- (7). Such equations in the problem of
focusing electron beams are considered in [32].
Thus, constructing variational equations requires
knowledge of forces and the second derivatives of the
potential energy
/<922n\ / d2n \ /d2n\
\du )0’ \dvd(3j 0 ’ U/3V o’
both being calculated on the supporting trajectory.
Let us also consider the case of the resisting force
depending upon the velocity
F* = —/ ( v * ) T*. (11.12.16)
By virtue of eq. (4)

. 1 /. ea
T* = —v* =r + n( i/H------- f)i+bl.+ }
Vp 1Ja
(11.12.17)
In the unperturbed motion

F = -f(v)T, F T = —f (v), Fn = 0, Fb — 0. (11.12.18)


Noticing that
642 11. Perturbation theory

we obtain
F* — FT = —f (a) ( e - y ),

* = -W+7-7 (11.12.19)

1 1 . 1 3 Perturbed Keplerian motion over a circular orbit

In the central force field

F = -^(r)£

(11.13.1)
the particle under certain initial conditions can move along a
r
n=— (11.13.2)
r
and thus
2
v
Fn = —, FT = 0, Fb = 0.
(11.13.3)
r
The second equation shows that the velocity of the particle
retains a constant value throughout this motion. Due to the
v = \/rty (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 = oo. We
obtain

r*=r + r(£ —y)+n^ + |^+b/3.


r
n(r*) = ip(x)dx. (11.13.6)
dr* i , ar* r
~dv = —r y* ^ dv = lT-.(-i
r*
T+n),
Vr/ y* ^
i , dr* 1
* K &Perturbed Keplerian motion over a circular orbit 643
11.13
O
C
*1

= —r
a/? y* ^ '~dp y% ^ y* ^
The second derivatives of this expression are required for
formulae (12.15). To this aim we notice that

Hence,
dll . , dr* r v... ip(r)
_=v,(r)_cs-_^(r) + _^W, W*—I3

and
/a2n\ V>(r) / Qr*
U^Vo - (?*)
\dvj o r ► (11.13.7)
/ a2n \ /a^n\ _Hj)
\dvdp)0 ’ \dp2) o r'
Taking into account eqs. (12.15), (3) and (7) we cast the
variational equations (12.5) as follows

rr
i>++„ {r = 0, (11.13.8)
/ (r) )
J+ = 0.
r
In particular, in the problem of perturbation of the circular orbit
of the earth satellite
<Mr) = 4 = TT’ =2/x
-3f2tr
= “TT

and equations (8) take the form


v
e — 2 —is = 0,
r
. v. 3v
2
v + 2-e------yu = •• (11.13.9)
••
v
or
^§-2^=0, p^+2^-3v = 0, ^+/3 = 0.
(11.13.10)
644 11. Perturbation theory

Here we introduced the new independent variable


u=-t,

(11.13.11)
r
which is the angle of rotation of the radius vector measured
from the ascending node of the circular orbit. The solution of
de
3 (2VQ — £Q) + 2 (2£Q — 3I/Q) COS U + 2v' sin u, 0
du
v— 41/0 — 2£Q + (2efQ — 3i/o) cos u + (11.13.12)
v'Q sin u,
(3 = (3 cos u + (3 sin u.
f

The first equation determines the perturbation of the orbital


0 0

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
^ . 2 on
• • «the supporting trajectory
<Pr = -a—- sin zsin2w,
, RW , 3 . 2 - 3.0.
= a-^r- 12u
— - sin i + - sin i cos (11.13.13)
Z Zd
F&v2 .
= —a 3 sin 21 sin u.
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
v
-r & r — = o3, a I = a—, R*
~FT ~ Cl, ——02,
Ho n 0 Ho r (11.13.14)

Robeing the earth radius, we obtain the following system of


equations
d261 0dS2 .2..
——z o—2—— = —O'! sin ism2u,
du du
d2S2+^dSi f3.2A3 .2..
~dv? ^~du ~ ^ Qi 1 1 ~ 2 Sm 1 + 2 1 ai
Sm 1 Sm
d 6s
2
+ 63 = — oq sin 2i sin u.
du2
(11.13.15)
11.13 Perturbed Keplerian motion over a circular
645 orbit

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 conditions has the form
1 3 . 2 A /\ 1.2A• 5.0..'
8 i = 2oq 1 — - sin 1 I u — I 1 — - sin 11 sin u + — sin
zsin2w
1 1
82 = —ot\ 1 sin i cos u — — sin2 i cos 2u — 1 H— sin2 i
83 — -OL\ sin 2i (u cos u — sin u).
(11.13.1
6)
The radius vector of the particle in the perturbed motion is
given by
h* = -n (l - ^62) + y (t8! + bfi3), (11.13.17)

which follows from eqs. (2), (14) and (12.3). The


corresponding unit vector e* is as follows

cr — —n 4--(T8 1 -f- M3).


-------------(11.13.18)
Let m, m1, k denote the unit vectors of the line of nodes for
the unperturbed orbit, the perpendicular to it in the equatorial
plane and the perpendicular to this plane, respectively. Then,
as Figs. 11.2a and 11.2b show
r = — m sin u + cos u (mi cos i + k
sin i), (11.13.19)
n = — m cos u — sin u (mi cos i + k
sin i),
and the expression for e* takes the form

1 ^0, .
= m ( cos u------0i sin u
r
1 • . . -^0 n • Bo c • • \ ,
mi cos 1 sin u H 81 cos 1 cos u- - -83 sin z +
k ( sinzsinix -|- --£1 cosixsinz
^0H ^<53 cosz ) . (11.1.3.20)
,
r
646 11. Perturbation theory

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 27T + x/j where x/j is small. In this position
cos 9 = e* • k = sin i sin u + —6± cos u sin i + —63 cos i = 0.
r r
Then we obtain
*fR01
—27T
sin 1 + — a 2(l — — sin2 i) sin i + sin i cos2=i 0
and
moreover
ip fRoY ( 5.2. (11.13.21)
2IT \ rJ \ 2
For this value of u the expression for e* is
2
* ^ . (R0 cos i[ 2 — “ sin2 i) — sin2 i cos i
er = m+27rmi < — cos 1 + ' a
27T
or
e* = m — 27rmi Ro acosz. (11.13.22)

Denoting the angular displacement of the node during one


revolution of the satellite by x we obtain

acosz. (11.13.23)
11.13 Perturbed Keplerian motion over a circular orbit 647

This coincides exactly with formula (9.14)


for e = 0. Let us introduce the variable x
defined as
'll)
follows
u=x (11.13.24)

It follows from the above said that cos#, being a periodic


function of u with period 2n + T/5 27r-periodic function of x.
18

Turning to eq. (20) we obtain

cos 9 = e* • k = sinsin
' Xl() Ro (Si cos u sin % + 6s cos i),
i x + — cos +
x
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
2 — — sin22 % ] cos x — — cos2 x sin2 i
cos 9 = sin i sin x + cos i
3 / 6
(11.13.25)
Denoting the longitude of the satellite by A we have
. n x * A = e* • m = cosrr —
sin#cos x'tp . Ro c / x •
—— smx----di (x) smx
/ 27r r
\
( x^ip \
sin 0 sin A = e* • mi = cos % ( sin x + —— cos x +
^V 2;r ;
— [61 (x) cos x sin i — 6s (x) sin
i\ r

or using the above formulae

L 0 COS A = COS X x tan x cos i+


2 — - sin % r
x .9 o . o . . o
------ -----sin x — - sm i sin x (11.13.26)
cosx 6
R20
sin 9 sin A = cos % sin x < 1a [x cot x-\-
2 — - sin22 i ) cos x — ^ sin2 i cos2 x
— sin i 3 /6

and therefore
f R?
tan A = cositanrr<l-------—a [x (cot x + tanrr cos2 i) +

2
— . (11.13.27)
-
si
648 11. Perturbation theory

Now we express variable x in terms of time t. Since


longitude A 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
Pa = r* sin2 6\ = (3 .
2
x

for the perturbed motion. However the perturbations ^1,^2, £3


and their derivatives are equal to zero at the initial time instant
which allows calculation of the constant f3x under motion on the
circular orbit with velocity
f3x = (r x tv) • k = —rv (n x r) • k = rv cos i.

Thus,
rv cos 1 (11.13.28)
A=
r*2 sin2 0
On the other hand, by differentiating expression (27)
we obtain x 11 — a [cos2 x + sin2 x cos2 i + 2x tan x cos2 i+
cos2 A
A cos i . f „ RQ
2 — - sin2 i 11
—------(l + sin2 x) —— sin2 i .
cosx

Expressing x, replacing A and sin 0 cos A by means of eqs.


(28) and (30) and taking into account, that due to eq. (17),
r*~ =rUl + 2^-a . 1.9A1.2• „ 1o
1 — - sm 1 cos x — — sm 1 cos 2x — 1 + —
(11.13.29
)
we obtain, after a little algebra, a very simple result
. v 1 1 ^0 (o ^ . 2 •
x = — 1 H—%-a 3 — - sin (11.13.30
r
)
From this equation we obtain

t — to = —x v
Rl
a [ 3 — — sim i
(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 = 2TT. Designating this time interval by T we
have
1 ^0 (o ^ . 2 • (11.13.32)
T = To 1
~ ~^2a ( 3 ~ 2 SU1 *
11.14 Equations for perturbed motion of a material system 649

with To being the period of the unperturbed motion on the


circular orbit
To = 27T—.
V

In passing we note that another formula for period T was

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 representative point on
Riemannian manifold Rn determined by the square of the arc
element
da2 = 2T(dtf = askdqsdqk. (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 equations of motion which can be set in the form
(7.8.12)

r+{^ = Qa = aa0Qp,
(11.14.3)
expressing equality of the contravariant components of
acceleration and force of the representative point. Equations
(2) can be considered as parametric equations of the
trajectory of the representative point, i.e. the supporting
trajectory. The unit vector r of the tangent to this trajectory is
determined by its contravariant components
daa 1
ra = — = -qa.
p = xarcn (11.14.5)
rabeing the basis vectors.
The equations for the perturbed motion are obtained by
replacing the generalised coordinates qa in eq. (3) by qa + xa
such thata
r+x { | (q0 + i0) (q7 + i7) = <2* + /**“, (11.14.6)
650 11. Perturbation theory

where Christoffel’s symbols and forces are taken for the


perturbed motion. Here p is small parameter and thus the
additional forces given by their contravariant components
<3>a are included in the variational equations of the
supporting trajectory. Up to the first power in perturbations we

(11.14.7)

with the forces being assumed to depend on the generalised


coordinates and velocities.
Finally we arrive at the variational system of equations
dQa dqP
xa + x°(f(p dq6
= x@
dqP + x
f (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 character of the quantities in eq. (8).
The vectors of velocity and acceleration are introduced into
consideration. Differentiating eq. (5) using rule (B.6.1) yields
^ > = r «( i ” +
{ /3 “ } ^)

and moreover (11.14.9)

P = rc ■fx' + { “7 u^7+
1 £ }'
q q~
0

{ £ } + (**+ { A }‘fx") {l K •
Replacing here q@ by means of equations (3) for the
unperturbed motion we have
{£ }*v+*w({ a a
P = rc xa +2 cr6 Pi
a
(T xM .
7
(11.14.1
The coefficients of ra in the expressions for p and p are 0)
the
contravariant
s|c Ot
components of these vectors. Let us denote them by x and x ,
(11.14.11)
11.14 Equations for perturbed motion of a material system 651

r= }-

{ °e }{ fiy })+2{ 0y }^ + { £ <U14-12>


* oc **0;
Now we replace xa and xa in eq. (8) by x and x . The principal
significance of this transformation is that we use the components of
the invariant quantities, namely the vectors of velocity and acceleration. If we
assume = 0 and the forces to depend only on coordinates,
then we come to the following equations

The expression
|^-+{;7p=w (ii.i4.i4)
is the covariant derivative of the contravariant components of
the force vector Q. Recalling definition (B.14.2) of the
Riemann-Christoffel tensor
p.Q;
n
^8f

(11.14.15)

we can rewrite variational equations in the form


xa xV - XsVsQa = 0 . (11.14.16)
Using Ricci’s theorem, see Sec. B.5, we can cast these
equations in the covariant components
<w xa +RlS0acPx6q0 - xsVsQa = 0. (11.14.17)

The covariant components of the Riemann-Christoffel tensor


are determined by formulae (B.14.5). They have the following
properties:
1) symmetry with respect to pairs of subscripts 76 and (3a
=
R^/8f3a Rj3cr^(8’)
2) skew-symmetry with respect to subscripts 7 and <5, (3
and a
R^f8{3<7 = =
R^8j3ai
(11.14.19)
652 11. Perturbation theory

3) Ricci’s identity
R1 8(3a + Rsp-fa + R/3~f6(7 = 0, (11.14.20)
expressing the property of cyclic symmetry with respect to the
three covariant subscripts.
In the case of potential forces equations (17) take the form

+RiS0a(pXs q13 = -xsVsn(T = -xl/ <9 n


2
x
\dq8dq°
(11.14.21)
The energy integral for the unperturbed motion is as follows
t + n = fc,
whereas that for the perturbed motion is
T* + II* = h + Sh,

where 6h denotes variation of the potential energy. In the

latter expression T* = \a*a0 (qa + ±a)

(q0 + x0) «T+ + aa qax


0
0

= T + \ ([a, T, 0\ + [0,7; a]) qaq0xl + aa0qax0.


Here we used the representation for the derivatives of in
terms of the Christoffel’s brackets derived in Sec. B.4. Taking
into account eq. (11) we obtain
T* - T = [a, 7; 0\ (fifx + aa qa xP -a j j x^ifif
1
0 a0
1

= [a, 7; 0\ cpifx1 + aa0qa xP -qaqa [cr, 7; a] xy = aa0qa x&

and the relationship expressing the variation of the energy


integral takes the form
.a *P on a (11.14.22)
aa0q x +—a; =Sh-

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 aa/3 and, second, the equations for the
unperturbed
motion (14.2) for n = 2. Basis vector ra having direction of the
tangent
to the coordinate line along which parameter qa changes is no
longer the
derivative of vector r belonging to the manifold with respect to
qa. Vector
r can be determined only on an Euclidean manifold which
includes the
Riemannian one under consideration.
The unit vector r 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)
t = cjn, uu = ka,

(11.15.1)
where k denotes the first curvature. Frenet’s second formula
(for n — 2)
U=-U;T

(11.15.2)
is easy to obtain by differentiating the relationship r • n = 0
and taking
into account that n has direction r (or opposite) so that n • n =
0. It is
worthwhile noticing that eqs. (1) and (2) are equivalent to
relationships of
the form *CK *CK *CK *CK
r = cun , n = —cur ; r = uona, n = — ura (11.15.3)
(11.15.6)
linking covariant and contravariant components of vectors r, n
654 11. Perturbation theory

aaa x ra —e — 20 cu — vcu — ecu2,


* n (11.15.7)
aaa x n° = v + 2ecu + ecu — vcu*.

Returning to the variational equations (14.21) and taking into


account that
x6 — er + vn6, q
6 1
= <TT7 , (11.15.8)

we come to another form of these equations


£ — 20cU — VCU — ecu2 + <J2R76l3a (^T7T6U° + VT17l671°) +
(er + vn6) raW Ua = 0, 6
6

(11.15.9)

v + 2ecu + ecu — vcu + & Rl pa (£T7T‘V^T°’ + ^T7T‘V^T°’)


2 2
8

+ (er + vn6) naV Ua = 0 6


8

K (a, c) = R d^ar a r^c°, 1


1 6
(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
J4
a#
*=1 Rjspa,
(11.15.12)
where EA7<5 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
A11 — i r 7^2323) A12 = — i?2331j A13 = — #2312,
m N N
A22 = -j—r i?3131,A = — #3112, 23
> (11.15.13)
|a| \a\
A33 = -j—7#1212,
m
where \a\ = |aa^|. 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 isRl212
A =------ 33
2 = K.
-----------(11.15.14)
&11&22 “ ^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
R"y6/3cr [ificr A- ^ • (11.15.15)
To prove this we make use of the following identity
(11.15.
16)
where 6* denotes Kronecker’s symbol. Substituting for Ricci’s
tensor in eq. (15) its expression (12) yields
\ RxpTUJ = \ (6*6% - Spp) {ST08% - 818%)
Rxpru>.
(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) = AXp £x l6 r^a6 €p0a T0C°.

Then, referring to representation (B.2.7) for the vector


products
rxa = p, rxc = q,

(11.15.18)
we
Let obtain
a = c = n,
where n denotes the unit vector of the normal to the curve.
Then
K (n,n) = AXpbxbp,

(11.15.21)
where b = r x n designates the unit vector which we refer to as
the second normal. Its covariant components are equal to
b\b)
K (b, b) = Ax>in\np, K (n, =Sa 76 Xli
= -A tV.bxnp = K (b, n).
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(r,n) = 0,
K(t,t)= 0.
(11.15.24)
The vectors r and n belonging to R2 have the components
T^T and n ,?? respectively (let us recall that kn is determined
2 1 2 a

as the covariant derivative of r ). Thus, unit vector b (which


a

does not belong to R2) has the single component


63 = y/H (t1??2 — T2in})
(11.15.25)
which is equal to unity because it is equal to the area of the
square constructed by the unit vectors r and n. Hence, for n =
2, due to eqs. (14) and (21)
K (n, n)
-R1212 (11.15.26)
K.
&11&22
0>\2
Let us return to eqs. (9) and (10). By virtue of eqs. (24) and
(26) they have the form
e— — VLU — ecu2 + (er 6
+ i/n6) raU$a = 0,
(11.15.27)
v + 2EUJ + eu — UUJ2 + &2VK + (er6 + vn6) n^Usa = 0,

where
d2U T1
TTZ = Has-
n<5<7 = = (11.15.28)
dq6dq a
\ Sa J dq T

The integral of energy, due to eqs. (14.22), (5) and (6), is


given by
att

& (e — ISLU) + (era + vna) = 8h. (11.15.29)

As follows from eq. (7.5.29) for the unperturbed motion


Using this result and differentiating
d=-|^r Q
, =eq. (29) we obtain the first
equation in (27). Then having the energy integral we can
remove it from consideration. The time-derivatives of ra and na
are determined by means of formulae
gu
(3) .
The time-derivative of the covariant component of the force m
the unperturbed motion is determined analogously by means
11.16 Systems with three degrees of freedom 657

Using eq. (30) we cast the energy integral in the form


be — be = 2bvu + 8h. (11.15.31)
Then, taking the derivative of the second equation (30) with
respect to time and using the first one we find
gu
glo + bio = ——ujT a

&r6naIlsa dqa
or
2glo + bib — — &T naIlsa-
6

(11.15.32)
Taking into account eqs. (31), (32) and (1) we write eq. (27)
as follows
v + v (b2K + 3b2k2 + n6nan^) + 2kSh = 0.

(11.15.33)
It contains only v and the parameters which are known from
the unperturbed motion. Under the assumption that v has
been found, the problem of determining £ from the energy

11.16 Systems with three degrees of freedom


We introduce into consideration unit vectors of the tangent r,
the first normal n and the second normal b. As before, n is
given by Frenet’s formula
(15.1) and b = r x n, so that r,n, b are determined at any point
of Rs and makes an orthonormal trihedron. The first Frenet’s
formula is now set in the form
T = V( i)Il, CJ(i) = fc(l)<7,
(11.16.1)
where fc(i) 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 • r = 0 we obtain
n • r = —r • n = ~^{i)
and taking into account that n • n = 0 we can write
n= + cj(2)b. (11.16.2)
658 11. Perturbation theory

Here factor LU(2) is introduced. By virtue of eq. (2) it is equal to


n • b = cj(2).

(11.16.3)
Keeping in mind that
b=rxn+Txh
and applying eqs. (1) and (2) we obtain the third Frenet
formula
b = cj(2)T x b = —cj(2)n
p = er + is n + (3b (11.16.5)
and obtain, instead of eqs. (15.6) and (15.7)
OL *

<W x Ta = k-uj(X)V,
OL * (11.16.6)
a
<W x n = i> + eu)(i)£ - 2)P,
OL .
*

<W x ba = (3 + U)(2)V.
Furthermore
X T° = £ - ch( 1)1/ - 2W(X)V - V(1}£ - W(l)W(2)/3,
GUrr X 71°
a;;
v + <h(i)£ — <h(2)/? + 2W(i)£ — 2U)(2)P ~ ( (i) "(2)
**<-* •• ry
aaa x b° = (3 + cj(2)^ + 2CC;(2)Z> + a;(i)a;(2)e^2)
- u3 ) 2
(3. (11.16.7)
Taking into account relationships (15.20) we can write
differential equations (14.21) under notation (15.11) in the
form
£ ~ )V - 2u)(i}is - Jfae - o^(i)C^(2)/3 + (er6 + isn6 + (3b6) T°Uas = 0,
(11.16.
8)

is + — cj(2)/^ “I- — 2uj(2}(3 — H- ^(2)^ v


a1 [K (n, n) is + K (n, b) (3] + (er6 + isn + nana<5 = 0,
6

(11.16.9)
11.16 Systems with three degrees of freedom 659

The energy integral (14.22) is


<7 (e - U(i)v) + (£TS + vn 6
+ /3b6) = 6h.
(11.16.11)

Equations for the unperturbed motion should be written


as follows
.. an Q 0„ =~^b (11.16.12)
r , CTLO. an
{1) =„-—n , an oq. a
a a
a= oqa
(/Qa
Differentiating the second equation with respect to time we
find with the
help of Frenet’s formulae and the two remaining equations
(12) that - 2-—TaW(i) + naTadUcrS = 0, uC[a (11.16.13)

whilst the third equation in (12) after the covariant


differentiating yields
acj(i)a;(2) + baTa&Iiao. = 0.

(11.16.14)
Equations (12) allows the energy integral (1) to be expressed
which does not differ from eq. (15.31).
Using now eqs. (15), (14) and (13) we can transform eqs.
(9) and (10), to get

i) + v a2K (n, n) + 3^^ — uo22) +


+ (3 [<J2K (n, b) - w(2) + Hs<Tn ba] - 2w{2)p + 2k(1)Sh = 0,
6
(11.16.16)

'$ + /3 [«J2K (b, b) - wf2) + ngabs


±v\K (

Quantity £ does not appear in these equations. It is


determined by quadrature 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 (n, n) and K (b, b) are
quadratic forms of na and ba, respectively, whereas K (n, 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.
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 cj( 2) = 0. These
equations take the form
v + P2V + q(3~ 2u;(2)/3 + 2k^)Sh = 0,
(11.17.1)
+ P2@ + QV + 2CJ(2)Z> = 0,
where
Pi = 3o^ 1} - u>^2) + (a2ASi3 + nSp) n6nP,
(&2aS0 + n^) b ^,
P2 = -^(2) + 6 (11.17.2)
q = (-a2Asp + lisp) b^n6,
Asp being the covariant components of Ricci’s tensor. The
integral of energy
(15.34) is
t

/ tSh
vdt T —;— T £Q.

(11.17.3)
lJe 2
+oi; +(3
2 2
(11.17.4)

remain bounded for any t however great it is.


The particular solution of the system of homogeneous
equations corresponding to system (1) is sought in the form

v = Neirt, (3 = Beirt. (11.17.5)


We obtain the system of homogeneous
(pi - r 2 ) N + ( q - 2uj{2)ir)equations
B = 0,
(11.17.6)
(q + 2uj(2)ir) N + (p2 - r2) B = 0.

The characteristic equation is


( P i - r 2) ( p 2- r 2) - q 2-4uj22)r2 = r4, - (pi +p2 +4CJ22)^ r2 + pxp2 ~ q2

= 0. (11.17.7)
11.18 Examples 661

Let us denote the roots of this biquadratic equation as r\


and r\. Provided that V\V2 ~ Q — 0, eq. (7) has a double zero
2

root and the solution of the system of inhomogeneous


equations (1) has terms proportional to t2. If r\ (or 7*2) are
negative or complex-valued then iv\ (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
Pi + P2 + 4W(2) > 0, (pi +P2 + 4W(2)) > 4 (P1P2 - q2) > 0
(11.17.8)
hold. This would be sufficient at 6h = 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
2kn)6h 2k(i)8h
2
(11.17.9)
P1P2 - q P1P2 - q2
and e remains finite under the following
condition P1P2 - q = 4uj q.
2 2
w

(11.17.10)
The frequencies of oscillations about the stationary motion
are 7*1 and 7*2. It would be an error to think that inequalities
(8) themselves ensure bounded values of v and (3 because
the variational equations (1) are derived under the
assumption that e is bounded.
A similar analysis of a simpler case of two degrees of
freedom leads, at 8h = 0, to the inequality
which, after Synge, gives the stability condition for the
stationary unperturbed motion. If 8h 0 the equality
Pi = 4 uJ or Ko + n san*nG — J = 0
2 2 2

should be added to inequality (11). The frequency of


oscillation about the stationary motion is equal to 2UJ.

11.18 Examples
11.18.1 Two particles attached together with a string
Two particles M\ and M2 of masses mi and m2, respectively,
are attached together with a weightless inextensible string of
length l. The string passes
662 11. Perturbation theory

FIGURE 11.3.
through an opening 0 in a smooth horizontal table. Particle
M\ 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 r —
OMi and p then the kinetic and potential energies of the
system are given by
T= ^ (mif*2 + m\r (p2 + m^r2) , II = —m2g (/ — r) = rri2gr + const.
2

(11.18.
1)
Constructing the differential equations of motion by means of
Lagrange’s equations we obtain
In what follows we study the stability of the stationary
motion
r — ro — m2^2 , Cp — (p 0 — const
(11.18.3)
mi^o
which is one of the particular solutions. Under this motion
particle M2 does not move whilst particle Mi moves on a circle
of radius ro with angular velocity p0. The string tension at any
point is m2g = rairo^o-
We notice that the motion of the system can be compared
with the motion of particle m = on the surface of a cone with
the vertical axis, with vertex O beneath and the half opening
of the cone a defined by

z — r cot a. (11.18.5)
11.18 Examples 663

Thus, the kinetic and potential energies of the particle are

T = -m [f2 (l + cot2 a) + r (p2] , II = mgz = m\gr cot a = rri2g*r


2

which coincides with expression (1) provided that g* = gy/rri2/mi.


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

an = mi + 77i2, ai2 = 0, a22 = mir2.


(11.18.6)
Equations of the unperturbed motion (14.2) have the form
Q1 = r0i q2 = <p0t.

(11.18.7)
The squared velocity a under this motion is 2

a = 2T = miro^o^
2

T1 = 0, =
da a (11.18.9)
The vector of the tangent is a unit vector since, due to eq. (8),
lap^T13 = a22T2T2 = = 1.
a

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
mir
(11.18.10)
mi +
77l2
Using formulae (15.3) we d a 2 ° + \pj\Ti3rl(7'
un
obtain
that is, in our case
"ni={22}^=-j miroVo cun2 = 0. (11.18.11)
(mi + m2) a
Value UJ — ka is determined from the condition that n is a
unit vector
aapnanP = annlnl = 1.
664 11. Perturbation theory

We find
/fj = JWo_ mi. 1
V'mi + m2 n =0.
2
a^mi + m2 ----;---- n
mi + m2 ’ (11.18.1
Calculation by formulae (15.26) and (B.14.5) yields2)K = 0
which we might expect since the Gaussian curvature of the
conical surface is equal to zero. Expression ILspnsnP is zero as
well because of

Equation (15.33) takes the form


Cl
i> + 3u2u + 2-----......-—
=0.
------------------(11.18.13)
?Wml + 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.
6h = 0) as equality (17.12) does not hold. For the problem in
question
2h = 2 (T + II) = (mi + m2) r2 + mir (p2 + 2m gr, 2
2

thus, applying (3) we obtain


Sh = (mir0</?o + m29) Sr + rmr^Cp^Cp = mx (2r0<pl8r + r$(p06cp) = ip06(3v
since r = 0 under the unperturbed motion. Here /? = mir^
denotes the
resultant angular momentum about axis Oz and, for this
reason, Sh = 0
for the perturbations which do not affect the angular
momentum. Such a
perturbation can be Xo
realised x1 =case
= (Sr)0: in the Xq =when particles M\ and
(11.18.14)
M get
2 Xo = (S<p) o , X r=0 ,
±1 2
= {6<fi) , 0

impulses directed along the string. 2_ Ina.2


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 — 0 x1
x1 = vn1 Z2 = £T2 =
y/mi + m2 ’
11.18 Examples 665

and thus
v = —yjm\ + ra2<5r, s = ^/mlro6(p.

(11.18.16)
Next, due to eqs. (14.11) and (15.6) we have

xl + jr2a:2<T = (^ + we) =
“ Wl/) r2>
which leads to
z> = —\Jm\ + m2 (<5r)* , £=
v m7ro<5(Jb,
/
(11.18.17)
£o = 2 v u>.
0

(11.18.18)
v = Uq cosy/3ujt H—^2— sin y/3uot,
y/3i (11.18.19)
e £o + UQ sin y/3ujt H—(l — cos y/3cut)
' 75 y/3c
It can be easily proved that condition (18) is met. Perturbation
of the polar radius Sr and the angular velocity S(p are
determined by eqs. (19), (17) and
(16)
Sr = (Sr) cos y/3cut + ^1° sin y/3uot,
0

y/3UJ
(<5r)0 cos y/3cot + -tf sin
r0S(p = —2(p0 y/3ujt v3 LU = —2 (f0Sr.

(11.18.20)
The angular momentum of the system about axis Oz 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 # = q1,\p — q — q we come to the following
2 3

expressions for the kinetic and potential energies


^#2 + 'll) sin2 #^ + C (^p + 'ip cos #^ , II = Qzc cos #
2

(11.18.21)

and thus
an = A, a\2 = 0 &13 = 0,
a22 = A sin2 # + C cos2 #, &23 = C cos #, f
(11.18.22)
&33 — C.
Resolving Lagrange’s equations for #,^,<£ we
have
•• C - A ■.2 . q C . • . .q (11.18.23)
$ H---- -—yj sm#cos# + —(/70sm# =——
ip +^ ^ ipi) cot $ —-7 . ■ = 0,
(11.18.24)
A A sin v

lb-----— (cos2 # + l^#+§^#cot # = 0,


-------(11.18.25)
sin# \ A )A

and then
in'#,
_C___
A sin # ’

cot#.
(11.18.26)
Under stationary motion of the regular precession the
angular velocities
of precession 'ip and spin Cp as well as the angle of nutation
# retain constant
values. Assuming 'ip = ipo,ip = = 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 i4^cos^° + = ~^p
{C(p0f >4(A-C) Qzc costio, (11.18.28)
11.18 Examples 667

eq. (27) has two real-valued roots


1 “F o
ipo =
2(A — C) cos 4 {A - C) Qzc cos 1}o ,
i?o (11.18.29)
defining slow and fast precessions.
For the perturbed motion
0 = + x, f + V-
We write equations for the perturbed motion in form (14.8)
where x1 = x, x2 = £, x3 = rj. Taking into account that
Christofell’s symbols depend only on g1 = # we find
9fll / 2 cos2l9
(j — A
cos
•C\
^ Y°~~ o + <Po^oj ^oJ ,

= 0 for 0 = 2 , 3 ,
since the non-zero Christoffel’s symbols with superscripts 2
and 3 have subscript 1 and $ = q = 0 for the unperturbed
1

motion.
We obtain the following equations

• 2 C—Ax + x •C Qzc
( —-— cos 2$0 ■Vo'&o-jcos
COS $0 +

+
+2{®iW=°-

Here it is taken into account that


<91 Qz
Qa = aalQi = -a' 1 c sint?,
W A
since a11 = A a12 = a13 = 0.
The two remaining equations are

l + 2{221}v’oi + 2{321}^x = 0,

»? + 2j231}^oi + 2j331jwi; = 0

or

C + 2|221|V’oa: + 2|321|v5oa; = <*,

71 + 2
\l^vx + = P,
668 11. Perturbation theory

where a and (3 are constants. They can be expressed in terms


of variations 6(3^ and 6(3^ of cyclic coordinates
dT _ dT
W d&

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 ChristoffePs symbols eq. (26) we arrive
at the system of equations
x + x Ul ^ cos 2tfo + cos
cos +
2 (cK — A) • C• C
sin $0 cos $0 + —V^r/sin^o + sin$o =
0,
A
/2A-C- q C . \ n
£ + x \ - - - —i p cot i 9 - A
0 0 ) = 0,
A A sin $o '
1 A - C COS + 1 ) ^0 + “7^0 ^0 C
rj + x
2 COt
= 0.
sin $o
(11.18.30)
Inserting the obtained expressions £ and r j into the first
equation in (30) and removing the angular velocity of the spin
( p by means of eq. (27) we obtain the differential equation
0

x + i p Q (l — 2/icos^o + M2) x — 0,
where
Qz (11.18.32)
c
Aii
Hence, the perturbed motion about the regular precession
is a harmonic oscillation of the angle of nutation with
frequency
U = '(JQ \J 1 — 2/i COS o + IX2. (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’Alembert’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 alternative is
variational statements dealing with the stationary properties of
certain values and enabling the complete replacement of the
above statements.
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

A. I. Lurie, Analytical Mechanics © Springer-


Verlag Berlin Heidelberg 2002
670 12. Variational principles in mechanics

of gravity of the sought-for curve must reach its lowest position


at equilibrium. This implies a variational formulation of the
problem in question. Let
,qn(t)

(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 of functions (1) is said to determine the true path of the
system whereas
any of the oon configurations
=
Qi (t) + 8qi, • • • 5 Qn = Qn (t) +
(t)
(12.1.2)
admitted by the constraints and infinitesimally close the true
path defines
the varied path. In eq. (2) variations 6qs imply arbitrary,
infinitesimal, dif-
ferentiable functions of time. dL c dL (12.1.4)
Let L denote the kinetic potential7which ;—fiQs +is7TT"
the difference
between the {uQs)
kinetic and potential energy. Along the true path
L ( q i , . . . , q n , q i , . . . ,qn,t) (12.1.3)
is a given functions of time. For a varied path the variation of
the kinetic Sq (t ) = 0, 6q (h) = 0 (s = 1 , . . . , n).
8 0 s

(12.1.5)
Let us now introduce quantity S referred to as Hamilton’s action
over
time interval ( t o H i ) and defined as follows (12.1.6)

S = J Ldt
to

Hamilton’s action is the functional determined by the set of


n functions
of time q s ( t ) . On the true path S takes a particular value S'*
and functions
q s ( t ) and q s ( t ) are the generalised coordinates and velocities
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 6q s and 6q s
we should replace L by L + 6L . Then, the increment in
Hamilton’s action calculated with the mentioned accuracy is
given by the equality ti
5* + 6S J (L + SLdt
SL)dt
and is called the variationtoin Hamilton’s action. According to eq. (4)
we obtain
dL c dL
6S = + ttt-/cv N(6q8) dt. (12.1.7)
dqs dqs '

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
n ti n ti
ss
=e/ (H -1H) ^=-g/& {L)6q‘it• <i2'i!,)
s S1
^ to to

12.2 The Hamilton-Ostrogradsky principle


Turning to expression (1.9) for variation in action we have two
possible lines of reasoning. First, from the position of
Newton’s axioms and D’Alembert’s principle we can say that
qs (t) satisfy Lagrange’s equations

£s(L) = 0 (s = l , . . . ,n)

(12.2.1)
under the real motion and thus
SS = 0.

(12.2.2)
672 12. Variational principles in mechanics

Given the functional of the sort (1.6)


Xi
J= J F (yi (x), , Vn (x) ,y[(x),... , y'n (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
Sys 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 t\ 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 6qs are
independent of each other we can adopt the following
bqi = 0 , . . . , 6qk-i = 0, %+1 = 0 , . . . , 6qn =

0,
where 6qk ^ 0. Equation (2) takes the form
(12.2.3)
to to
Assume now that
d 3L dL Q dtdqk dqk
£k{L) T
(12.2.4)

at t = t*. Then we can find such a time interval (t*o, t * i ) that it


contains t* and £k (L) retains its sign. But 6qk 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 <
f*o and t\ < t * i - Then,
£l t*i
J £k (L) 6qkdt = J £k (L) 8qkdt.
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,
£k (L) - o.
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
12.2 The Hamilton-Ostrogradsky principle 673

his papers on dynamics and optics published in the mid-


thirties of the nineteenth 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 papers 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
l=
J2 q^sdL
8
^ ~h = - H- (12.2.5)
6=1
Then
' (X dH
x
Ps (oqa) - — oqs
(12.2.6)
s=1 . dqs
Equalities
■ aH
i, *
q =
- w, <a = 1
""") (12.2.7)
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
61 = Y, ‘ ,, dH
Ps {oqs) ~ 75—oqs (12.2.8)
6—1 L dqs
Taking into account eq. (1.5) we have
tl
JPs (6qs)* dt = PsSqs\\l0 to dt

and therefore
11
6qsdt = 0. (12.2.9)
to
By repeating the derivation which yields Lagrange’s
equations from eq. (2),
we arrive at the second set of the Hamiltonian canonical
equations (s = 1,... ,n), (12.2.10)
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 us and
variations of quasicoordinates 8ns 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
^=E dT c , d(T-U)s_
—- -6<JJS -I-------
s=1 L -----07TS

Replacing here 8uos in terms of (8TT )* by eq. (1.8.3) and taking


S

into account the rule ”6d = d8” and property (1.9.2) of the three-
index symbols, we have
8UJS = (87rs)* + EE 7>t<S7Tr (12.2.11)
t=1 r=l

and moreover
dT d(T-n) A-A _ dT
SL
= Z&J:(s^' + E -WJ + EE*ar"- 6ns.
s= 1 s= 1 t=1 r=1

Integrating by parts

fx d dT
J

and noticing that variations of the quasi-coordinates reduce to


zero
<57TS(£O) = 0, Stts (t\) = 0 (s = l, . . . , n )

IV d dT _ dr d(T n)
8S = dt
+EE^«*+ 87rs = 0.
dt dtos t=1' r=1 dirs
/ E
(12.2.13)
Since quantities 8TT are independent of each other we come
S

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 approach
allows derivation of equations of motion for the free body by
means of the Hamilton-Ostrogradsky principle. The derivation
12.2 The Hamilton-Ostrogradsky principle 675

motion of a rigid body based on relationships between


variations of projections 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
11
SS = j S'Wdt = 6S + QsSqsdt — 0. (12.2.14)
to
As follows from the above
11
d d L
£s (L) Sqsdt Sqsdt,
dt dqs
to
(12.2.15)
and equality (14) reduces to the form

dt dqs s
Sqsdt = 0, (12.2.16)

which yields Lagrange’s equations. If we introduce variations


of the quasicoordinates and represent the expression for the
elementary work in the form of eq. (5.1.6)
n
S'W = 'Y^PSSTTS, (12.2.17)
S= 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 S 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
S'R = SS +/ S'Wdt = SU + S'W) dt (12.2.18)

turns to zero, however no functional R exists since there is no


value whose variation equals S'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
(xi,... ,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 d2f and, in some exceptional
cases, by means of the sign of higher-order terms. Thus, the
condition SS = 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 S S, the minimum occurring
2

under the condition S2S > 0. Limiting our consideration to sta-


tionary constraints it is easy to show the presence of a
minimum of the action S for a sufficiently small time interval t\
— to. To this end, we construct the expression for the

S=1 k=1

Here the calculation is limited by terms of second order in 8qs


and 8qs. 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

(12.3.1)
12.3 On the character of extremum of Hamilton’s action 677

Taking into account that 6qs (to) =0 we


have
ti

where (3S denotes the maximum of the absolute value of 6qs


for to < t < t\. Hence, for a sufficiently small time interval t\ —
to the term T (6q) prevails and determines the sign of eq. (1)
2
8 L « T (6q) >0,
since the quadratic form T is positive definite. Thus

(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 Aq and A\ given by
the generalised coordinates qQs at t = to and q\ at t = £i,
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 Aq and A\ 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
678 12. Variational principles in mechanics

FIGURE 12.1.
for the first time, see Fig. 12.1. The neighbouring path AqHA\
is assumed to intersect the true path AqBA\ at instants t = to
and t = t\. The action along the true path AqBA\ is denoted by
S*. Along any neighbouring path intersecting the true one at
points Ao and A\ the action differs from S* by a second-order
value. According to the formulated problem, the action along
the neighbouring path AoHA\ is equal to S* with accuracy up
to second-order values since S2S = 0 along it. Let us show
that this neighbouring path is the true path. Let us suppose
the opposite, that is suppose that path AoHA\ 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
AoERFA\ is smaller than along path AoEHFA\ and in turn
smaller than S*. But this contradicts the condition that A\ is
the first position on the true path AoBA\ where the second
variation S2S becomes zero along the neighbouring paths
intersecting the true path.
Therefore, positions Ao and A\ 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 A\ be the conjugate kinetic foci. We consider
the true path ABAiQ whose final position Q is reached after
the focus A\ has already been passed, see Fig. 12.2. The
action along this path is no longer the minimum. It follows
Sntq < SNAi + *SUiQ-
Thus
S'* — SNTq
— SabaiQ = Sabax + Saxq = Saha1 + S^q
= Sahn + Snai + Saxq > Sahn + Sntq = Sahntq>
12.3 On the character of extremum of Hamilton’s action 679

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 provide 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
n variables
L = ^PsQs,qn,p!,... ,pn)
S= 1

and replacing qs,ps by qs + 6qs,ps + 5ps we obtain


s2
n ^ n n dqsdqk
l = E^-oEE 8qs8qk +
s=12 s=1 k=1
o d H K K dp + -—-—opsdpk
cf
2———oq k s (12.3.3)
dqkdps dpsdpk
where the second derivatives of the Hamiltonian function H
are calculated on the true path. For the quadratic form in
expression (3) for S2L we adopt the notation
^nn
(aq, • • • , xn, ui,... , Un) = — ^ ' 'y dxsdxk xsxk+
'
l k=\
n d H
2
d2H 1 71 dfl, dQ .
2 ^ ^ XkUs ^vs: I, +
as;“ ,
i’ < -->
l2 3 4
dxkU
dusUsk
2
s= 1

dusdunk
^-sE an an 1
d^ .
bPs Sqs - d8 - t>qs sps + d6q
s=1 ps s S= 1
(12.3.5)
680 12. Variational principles in mechanics

Then, taking into account the condition for the neighbouring


paths
6 qs (t0) = 0, Sqs (h) = 0, (s = 1,... , n),

(12.3.6)
6 S ■i/5
2
6ps ( 6q: / . an
c c
dt. (12.3.7)
d6p, - oqs ops + d6qs

In what follows we return to this expression. Now we


consider the equations of motion which are infinitesimally
close to the motion under consideration, i.e. the variational
equations for the canonical system of equations
dH . OH . ,
. P'-sj 7 (» = i- (12.3.8)
dps
Constructing these equations in the way explained in Sec.
11.10 we obtain
. dSi . dSl , , ,
x =
(12.3.
‘ *r,' - ~w. <*“*•••••">•
u =
9)
where xs and us denote variations of coordinates and
momenta, and VL is the above quadratic form (4). Also we
notice that for Lagrange’s equations
dL . dL
Ps = 7T-, Ps = (s = l,...,n), (12.3.10)
dqs dqs

the variational system of equations


(12.3.11)
will be
OF . dF
= us = -— (s = 1,... , n).
dxs, x ) — ^ ^ ^ ^ dxs a 2L
F (xi,... , xn, X\,... n 's K . XkXs~\~
2
\dqsd k^ ~
q
'
~dq k dq s
d2L ^nn
. %S'Ek-sEE (p'sk%s%k F ‘Fbsk&sXk F Csk%s&k) •
dqsdqk s=1 fc=l (12.3.12)

As seen from formula (1) the second variation is


represented by the same form F of variables 6qk and 6qk, so
that
cl ••• C
? ... *7l

& ••• c? Vn • * * <


C”
cl ...
Cn C” ^ ... n
u

=1

and thus solutions (15) are linearly independent. According to


the definition
of the integral of the system of equations we have
C(to)=6sm, v?(to)=0, C(*o)=0, K(to)=6sm. (12.3.16)
We also need equality (10.7.8) for the Lagrange brackets
[0 s ’ 0n E ' 9qk dpk dqk dpk = X>*C-«) = °-
fc= 90s dpm 9Pm 90s
1
Having solution (15) and using its property (16) we can cast
the integral of Cauchy’s problem for the variational system of
differential equations
n
Xs = ^2 [X™ (<o) £? + um
(to) v™ (s = 1,... , n) (12.3.17)
m=1n
Us = tXm ^ + Um (fo) K
ra= 1
Let us return now to expression (7) for the second variation
of action. It is equal to zero if the equalities
dsi dn (12.3.18)
96ps $qs = dps = - 96q (s= 1,... ,n).
s
dqi dq,
7?} . " dPn

Vn • •• < dqn dqn


dp, " " dPn

(12.3.20)

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
6q.(t1)=J2u™(to)ri?(h) = 0 (s = l,...,n), (12.3.21)
m=1
which has non-trivial solutions for variations of momenta (£q)
provided that
A(*i,*0) = 0.
(12.3.22)
Here t\ denotes the root of equation A(t,to) = 0 that is nearest
to founder condition (22) the unknown parameters (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 t\ at conjugate kinetic foci whose
positions are given by the formulae
Qs (to) = Qs (tl) — Qs (t to, ai, • • • 5 i • • • 5 0n)
12.3 On the character of extremum of Hamilton’s action 683

It remains to prove that for any t from interval (£(Mi) when


the inequality
A (Mo) ^0 (t <t< tx),
0

(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 S2S 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
Ynn
^EE< sk6qs6qk = T (6q) > 0,

(12.3.25)
s=l k= 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
Id n
2jt^2Kk8qs6qk, (12.3.27)
S=1
to the integrand in eq. (13). Because the latter quadratic form
is a complete 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
1 K n n
S2S = /EE ask8qs8qk + 2bskvs6qk + 2bsk8qk ^ 7sr8qr +
to S=1 k=1
r=1

Csk'Vs'Vk T Csk EE IsrlkhSqMh + 2cskvs^lkr8qr +


r— 1 h=l
Xsk6qs6qk + 2A sk6qk ^ 7 sr8qr dt (12.3.28)
r= 1
and our goal is to choose such functions of time 7sr and \sk that
this expression becomes
.y K n n K
S2S -5 EE CskVsVkdt — / T (u) dt > 0, (12.3.29)
to S=1 fe=1 to
684 12. Variational principles in mechanics

where, by virtue of eq. (25), the equality takes place only for v s
= 0 for all s = 1,... , n. Hence, it is necessary to convince
ourselves that all 6qs are zero, that is S2S = 0 is possible only
along the true path (14). Hence, it is necessary to satisfy the
conditions
^sk “1“ ^ ^ Csr^frk "b A sk — 0?
r=1
CLsk + 2 ^ ^ (brk + Ark) Ifrs T EE Crh'lrs'lhk T > (12.3.30)
Ask —r=
01
r—1 h= 1
(s,k = 1,... ,n), J
where the second set of conditions is transformed by means
of the first set to the form
nn
&sk — A sk + EE Crhlrslhk (s, k = 1, . . . ,Tl) . (12.3.31)
r=1 h=1
The matrix of coefficients Xsk of the quadratic form (27) is
symmetric however bsk ^ Aks- Thus, the equations
n
^>sfc + ^ cfer7rs + Afes = 0 (k, s =
1,... ,n) (12.3.32)
r= 1
should be added to the equations. It remains to show, first,
that all equations obtained have solutions and, second, that
all 6qs are zero at vs = 0. Notice that, due to eq. (26),
n
qs ='Y^lsMr (s = 1, • • • ,n)
6

(12.3.33)
at vs = 0.
Let us return to some properties of the solutions (15) of the
variational
equations. Let us assume that
n n
il? = '529akrik, K = ~J2lskrlk {s,m= 1,... ,n), (12.3.34)
k=1 k=1
where gsk and lsk are functions of time which can be found
from the above
relationship for any fixed s under condition (24). Matrix l sk is
symmetric
which can be proved by means of eq. (10.7.8)
\Ps,Pm]
= E - «) = - E K E- c E «
k= 1 k= 1 \ r=1 r=1
nn
= EE^(lrf“W=ft
12.3 On the character of extremum of Hamilton’s action 685

Applying condition (24) twice, we obtain


Ikr = Ikr {k>i V — 1, . . . , Tl) .
(12.3.35)
Let us now relate functions gks and lks with the coefficients
of the differential equations (11). Substituting the particular
solutions and rf? into the first set of equations (11) for u s and xs
we have
n n n/ n
- 12 ?k = 12 ( i™+csknT) = 12 [
lskr bski bskT]k + Csk
12 ™
9krTi

k=1 k=bsk1 k=l \ r=l


12 + + 12 Csr9rk fVf =0 (s, m = 1,... , n).
k=1 \ r= 1
Returning to eq. (24) we
find
^sk “1“ &sk “1“ ^ ^ CsrQrk — 0, r= n1
Iks “1“ &ks “1“ ^ ^ Ckrdrs 0 =
(^, k (12.3.36)
= 1, . . . , Tl)
r= 1
From relationships (34) and the second set of equations (11)
we obtain
m = n (. n
\ n
( n
\
~ ( IskVk + Isk 9krV™ J = | UskV™ + bks gkrV™ ] •
k=1 V r= 1 / k=1 \ r= 1 /
Equating the coefficients of g™ and using eq. (36) we arrive at
the equalities
nn
Isk “1“ &sk EE crh9rs9hk — 0 (-S, k —
r= 1 /i=l
Comparing eqs. (36), (37) with eqs. (30), (32) and (31) we find
that the latter can be satisfied by taking
2fsk 9ski ^sk Isk (^5 k — 1, . . . , 77,)
(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 62S to the form (29).
Equations (33) (12.3.39)
which are now cast in the form
n
dqi dqi
dpi " ' 9Pn

dqn dqn
dp\ '' " dPn

(12.3.44)

is identically equal to zero at t = 0 since equations (42) are


identically fulfilled for arbitrary momenta p°s at this time
instant.
12.3 On the character of extremum of Hamilton’s action 687

When the initial momenta are p°s+8p°s, with <5p° being


arbitrary infinitesimal values, we have
Qs (h,Pi + f>Pi, ■ ■ ■ ,Pn + Sp^) — qs + dqs +■■■+ +•
dp\ dpi
(12.3.45)
where dots denote terms of second order of smallness in 8pk
and higher. If Jacobian (44) at t = t\ (different of t = to = 0)
equals zero again, then the system of equations
dp{{ +
dli6p°n °
(s = 1,. ,n) (12.3.46)

will have the non-zero solutions 8p°k. Then, due to eqs. (45)
and (42) we obtain
Qs (tl. Pi + <5^1, • • • ,Pn + SPn) ~ Qs (tl,Pi, • • • iPn) + • • • (s
= 1, . . . ,7i)
(12.3.47)
Thus, if at t = ti Jacobian (44) is equal to zero, then there
are two systems of the initial values of momenta
Pk and Pk+Sp°k,
which result in two infinitesimally close paths, namely path
(42) and the path

9»(*,Pi+tfPi,... ,P° +8pl) (s = 1,... ,ra),


both bringing the system to position Ai or to a position which
differs from Ai by small values of second order or higher. In
this case Ai 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 on equator (i? 0 =
7r/2, Ao = 0). Let us denote the projections of the initial
velocity on the directions of the tangent to the meridian and
equator by $o and Ao. Then the particle moves along the
great circle inclined to the equatorial plane by angle z, where
. Ao . . $ •=o
cos i — —, sm i —----- + \ 0>
u)
u> denoting the constant angular velocity. The arc along the
great circle is u>t and, as follows from the spherical triangle
AMQ, we have
688 12. Variational principles in mechanics

FIGURE 12.3.
see Fig. 12.3.
Therefore 1 / • 2 -2 \ — sinutcos+
sin$- m_ _ utj ,
CO'
d#0 ~
^oA
sin$- dd _ o (— sin ut + ut cos ,
3A0 '
1 d\ _ ^qA art
- tana;/; +
cos2 A d$0 q COS 6 Ut
2

1 <9A _ 1CJ/ /-2


;2
cos2 A d\ 0 6J0 COS 6 ut
2

Jacobian (44) is
equal to
A (t) = ^7 ; sin ut
uz
and becomes zero at t = it/u and A = 7r, which correspond
respectively to the position $1 = 7t/2 and point F diametrically
opposite to point A.
As a second example let us consider the problem of free
vibration of the oscillator.
a The motion
cos ut is given by the equality
+ — sin
Q ut, (12.3.48)
U
where a and (3 denote the initial values of the generalised
coordinate and velocity, respectively, and u is the natural
frequency. Due to eq. (44) we
12.3 On the character of extremum of Hamilton’s action 689

have

(12.3.49)

that is the kinetic focus is reached at time instant t\ — tt/lj at


position qi = —a. Indeed, the expression

q = a cos ut -\- {(3 + 6(3) sin ut, (12.3.50)

with 6(3 denoting an infinitesimal value, is the solution of the


differential equation for the oscillator and represents the true
path which is infinitesimally 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) describes a
bundle of extremals emanating from point qo = a 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 = tt/lj + t. Let
us
by construct the with
the true path neighbouring path of
the final point from
thetwo parts: the
prescribed first
true path
(48), so that this part is passed in time interval 2r. Hamilton’s
action along the second part is a minimum if 2r < tt/uo. Then
denoting the generalised coordinate along the neighbouring
path by q* (t) we have
(12.3.51)

ql (t) being given by eq. (50) within the above time interval.
The positions of the final points of path q{ (t) and the true path
(48) are given by the equalities

Q = —a cos cur---(3 sincjr = q% (12.3.53)

Function q% (t) is the solution of the differential equation of

vibration

Q2 = Ci cos ut + C2 sin ut,


690 12. Variational principles in mechanics

where constants C\ and C2 should be determined by means of


conditions (52) and (53). We arrive at the system of equations
C\ cos ujt — C2 sin ujt — a: cos ujt-(/3 + 8(5) sinur,
uj
C\ cos ujt + C2 sin ujt = a cos ujt-\/?sin ujt,
UJ
from which it follows that

Ci = ol - tan ujt, C2 = — (f) + \-6/3


2 uj uj V 2

where 2ujt < it. Hence,

#2 (t) = Ta — ^ tancjr j coscjt + — f/? + ) sin


ujt,
fir it \ (12.3.54)
----r<t <- + t).
Vcj uj J
It remains to calculate Hamilton’s actions along the true and
neighbouring paths
7T
- +T
U
S
J
=\ J (<?2 - ^ V) dt,
IT
— +T
UJ
S* = S*1+S*2 = ± J (qf-u; qf)dt + ± J (qf-cu2qf)dt.
2

UJ
(12.3.55)
To simplify this calculation we notice that the action during the
time interval (to, ti) for the harmonic vibration

q = A cos ujt -1
-----------------H sin
is equal to
If 1
S = - (q2 - w2q2) dt = — sinw (i, - i0) x
to
[(B2-u2A2) cos a; (to +ti) — 2ABu> sin w (to +
11)] . (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

S= \(/32 — a2uo72) cos uor — 2a(3u sinJ lot] , S* = S — tan lot.


2uj L v 4 uj
S-S* = (8/3) tsanjT > 0,
2
(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
qs = q8{t,C!,... ,c2n) (5 = 1,... ,n),
(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 ci,... , C2 n. A set of
true paths C' infinitesimally close to Co corresponds to values
c& 4- 8ck of these constants. Along these paths
2n ~
6qs = ^2 fr7SCr (s = 1- • • • , n).
If paths C' intersects Co at instants t = to and t = t\ then the
corresponding configurations of the system qi°^ and ijS, 1"* are
the conjugate kinetic foci. We come to the homogeneous
system of 2n linear equations
0 (s = 1,... ,n),

having non-trivial solution for 6cr if the system’s determinant vanishes, i.e.
dQi\ ( dqi \
9c1)1 ■" \^c2n/i
dqn\ f dqn \
A (i,*o) ^Cl/l \dc2n)i (12.3.59)
dgi\ / dqi \
dci)0 \dc2n)o
dqn \ / dqn \
/q \ dc2n ) o
The time instant ti 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

^asfe9fc+as=0 (s = l,...,l),

(12.4.1)
k=1
then the variations describing the neighbouring path should
satisfy, in addition to eq. (1.5), the following equations
n
^2askSqk= 0 (s = l,... ,1). (12.4.2)
k=1
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 equations (1) should
hold also for varied generalised coordinates along the neigh-
^Jask (Qi + • • • ,Qn + Sqn,t) [qk + 6qk\ +
fc=i
a8 (qi + Sqi,... , qn + 6qn; t) = 0 (s = 1,... , t) (12.4.3)
must hold. Taking into account eq. (1) and restricting the
analysis to the first-order terms we find
V <-jf^L(Ikbqr + ask6qk + '£tts* = 0- (12.4.4)
Ek=1 Lr=l dCIr f^dqr
On the other hand, differentiating equations (2)
we have
n n r^^qrSqk + ask (Sqk)*
EE Oqr ~b ^ dadt~^qr ~ (12.4.5)
k=1 r=lL r= 1
Subtracting eq. (5) from (4) and using the rule ” dS = Sd” we
arrive at the equations
/ 9a s 9as
EE(^-^)^ +E(g-^l^«,

(s = 1,... ,/),
(12.4.6)
which are referred to as the conditions for the kinematic
feasibility of adjacent motion, see [77]. From a class of
variations Sqk 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
ddsk dasr das dasr
dt (r,k = 1,... ,n;s = 1,... ,1) (12.4.7)
dqr dqk ’ dqr
equations (6) are fulfilled automatically. Hence, any adjacent
configuration is feasible under an adjacent motion provided
that the constraints are holonomic.
Let us turn to the relationship
ru £
J2 k(L)Sqkdt = 0 (12.4.8)
to fe=1

expressing the Hamilton-Ostrogradsky principle. There is no


need to require 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 compares 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 8q\,... , 6qi in terms of the remaining in-
dependent variables 8qi+i,... , 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
nr l
/ m £k +
X! dt = 0. (12.4.9)
k=1 L s—1 .
Then we choose the multipliers Ai,... , A/ 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 i of motion in the form of (7.1.6)
&k (-^) + ^ ^ ^s^sk — 0 (^ — 1, • • (12.4.10)
• , Tl) ,
s—1
694 12. Variational principles in mechanics

which should be considered together with the constraint


equations (1).
On introducing quasi-velocities cus and casting the
equations for the non- holonomic constraints in the form
n
UJS = '^2askqk =0 (s = l,...,Z)
(12.4.11)
fc=i
we can set conditions (2) imposed on the variations of quasi-
n
= ^2 askfiqk = 0 (s = 1,...,/).
(12.4.12)
k=1
In eq. (2.13) the terms vanish if their subscripts s and t take
values 1,... , Z. No condition is imposed on variations 6tt s for s
= / +1,... , n and, hence, the coefficients of these variations
can be equatednton zero. This leads to the equations of motion
d dT + ’fi dT dT an
dt t=l+1 r=l du>r UJt dns dll (s = 1 + 1,... , n).
duos s
(12.4.13)

The following kinematic relationships


n
Qs = ^ ^ bsk^k (5 = 1,... , Tl)
(12.4.14)
should be added to the above equations. It is significant that
the constraints should not be taken into account when
constructing the expression for
dT
the kinetic energy since equations (13) contains terms with ——
for r =
OUJr
1,... , l also. It is also an error if we take into account the
presence of the constraint and remove l generalised velocities
q\,... ,qi 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 6qk 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
12.4 Application to non-holonomic systems
695

problem
11
S = J Ldt (12.4.15)
to
subject to constraint (1), [34].
In order to solve the problem we introduce the
function
l /n
$ — L + fls I CtskQk + S—1 (12.4.16)
\k=1
with /is being the undetermined multipliers, and construct
equations for the extremals for the functional
ti
R = J <f>dt. (12.4.17)
to
The variation of this integral is
/X>. £«■ given by
kQk +as\dt
8=1 V fe=1
to

11
l
y] (qkdask + CLskdqk) +dt<5&s
-(-
Ms
S=1 U=1
Kl /n
kSqk + as dt. (12.4.18)
5=1
to ^fc=1

The second integral on the right hand side vanishes by virtue


of the constraint equations (1). Integrating by parts and taking
into account the condition for choice of the neighbouring paths
(1.5) we have
ln nL
j'22'22 Vsaskdqkdt = - ^2 ^2 (f s sk + Hgdsk) Sqk
x a

k=1
dt.
=1 lk=1
8=1 8=1
to to

The right hand side of eq. (18) is cast as follows


ll
In r
ll
l
- '^2^2lisask6qkdt+ ^ t0 fe=i ^ ^ (qk^tisk dsk6qk) T <5us
dt.
to s=l lk=l
8=1

(12.4.19)
696 12. Variational principles in mechanics

After a little algebra


n
^ ^ (Qk^Q'sk Q'sk^Q.k)
“1“ Sas k=1
dasr qtSqr
\.^ da dct
-a^) + s gf (12.4.20)
dqr ~df
This expression becomes zero if we take into account the
conditions of the kinematic feasibility of the adjacent motion
(6). Denoting jis as As we obtain
11/» /* l
SR 6$dt = 6Ldt -
/ to to S=1 k=1
ti r
it p
£k (L) + CLsk^s Sqkdt = 0.
E/
k=1
t0 s=l
We arrived at equality (9) which implies that we have obtained
the equations of motion (10). Hence, these equations are
obtained from the constrained 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 feasibility 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 a\x example
+ a^y + aof the motion of a material
3z = a • v = 0,

(12.4.21)
where the functions of the coordinates ai,a2,a3 are
considered as projections 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
rot Ma — M rot a — grad M x a = 0.
The integrability condition obtained by removing M is
da da da da da
a- rot a = ai s ,2 + a2 i 3 + ^3 2
dy dz Hz dx dx (12.4.22)
da i da2 da3 2d i da2 + da
dx dx dx a
dx dx idy
da\ da2 da3 1 da i i da2 9d 2
dy dy dy “2 dy dx a
dy _
da i da2 da3 dai i da3 da2 1 da3
dz dz dz dz "h dx dz dy
n da2 dai da3 dai
u dx dy dx dz
1 da da2 n da3 da2
2 idy dx u dy dz
dai das da2 da3 n
dz dx dz dy

= def a + 0,

Here def a and 0 denote the symmetric and skew-symmetric


parts of tensor Va. Equation (23) is now cast in the form
v • 5a — a • Sr = (v • def a • 5r — 5r • def a • v) + (v • 0 • 5r—5r •
0 • v) = 0. The first parenthesis vanish since for any symmetric
tensor
a • Q • b = b • Q • a.

The expression in the secondv xparenthesis can


^ rot a j • Sr— f 5rbex written
^ rot a as
v • 0 • 5r—5r • 0 • v =

v • (rot a x 5r),
thus, the conditions of the kinematic feasibility of motion (23) is
expressed by the equality
v • (rota x 5r) = 0.

(12.4.24)
a 5r = 0,
698 12. Variational principles in mechanics

which enables us to represent Sr as the vector product


Sr = 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
v • (rot a x Sr) = v • [rot a x (a x b)]
= v • ab • rot a — v • ba • rot a = —v • ba • rot a
=0.
If we choose vector b in such a way that it is not perpendicular
to v, then
a • rota = 0,
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 /, see Fig. 12.4. End O 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 C'
of the chain (for a fixed t). Projections of the displacement of
point iV, i.e. mass m, are equal to u (/, t) and v (/, t).
The condition of inextensibility of the chain results in a
certain relationship between functions u and v which is a
x = a + u (a, t), y = v(a,t). (12.5.1)
12.5 Equations of motion of distributed systems
699

lr
X
FIGURE 12.4.

This is the parametric equation of curve C' for a fixed t. The


differential of arc da is da, so that

2
da + = da.

This yields

2 du 1 d^ = / du \ 2 / dv \
+ (12.5.2)
~2 \da J \ da )

The analysis is limited to small deflections of the chain from


the vertical, i.e. we adopt that a, v and their derivatives with
respect to a and t are small. Let a prime and a dot denote
derivatives with respect to a and £, respectively. Then if v,v f,v
are considered as quantities of the first order of smallness,
then, by virtue of eq. (2) a, a', a are second-order quantities.
Since (a')2 is of the fourth order of smallness we can neglect
it in eq. (2) and write eq. (2) as follows
u/ (12.5.3)
700 12. Variational principles in mechanics

The kinetic energy of the system to the above


accuracy is z
T = ^p J (u2 + v2) da + [u2 (l, t) + f;2 (i, t)]
o
z
(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
11 = —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 (a + u)dcr = J ada + J uda = + J uda, XN = l + u (l, t).
lx.

Let us remove u from these expressions by means of the


constraint equation. Taking into account that u (0, t) — 0 we
obtain
u (/, t) = J u' (a,t) da = — i J v'2da,
o o
z z i i i
J uda = au\ 0—J au!da — — J v' da+^ J av da = —^ J (l — a)t>'
l f

da.
0 0 0 0 0
Omitting the immaterial constant we find
z z z
II = ^pp J (l - a) v' da + ^rap J v' da = ^pp J (l + - a) v'2da,
2

o oo
(12.5.5)
where
171
;;——
Zi = pi
and p denotes the ratio of the end mass to the
chain mass. The kinetic potential L takes the
form
L = HI v2 — g (l + l\ — a) v' da + liv2 (l, t) (12.5.6)
12.5 Equations of motion of distributed systems 701

Now the variational problem derived from the Hamilton-


Ostrogradsky principle is stated as follows: among all
continuous functions v(a, t) having continuous derivatives with
respect to a and t for 0 < a < l and t > 0 and satisfying the
boundary condition
i;(0,*) = 0,

t\ t\ (i
v2 — g (l + h — a)
v'
to to yo
(12.5.
8)
stationary. Here the ’’true path” is determined by the sought-
for function v (a, t) whereas the neighbouring path by v+8v (a,
t). The variation 8v (a, t) means an arbitrary continuous
function in the domain of definition 0 < a < l and t > 0 which
8v (a, to) = 0, 8v (a, t\) = 0, ^(0,^
=0 (12.5.9)
and has continuous derivatives 6vf and 8v. The first and the
second equalities 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 O is immovable.
11 l
8S =p J dt J [v8v — g + h — a) v'Sv'] da + hv (/, t) 8v (/, t)
to (l 0
Next, by means of integration by parts it is possible to remove
derivatives 8v' and 8v. To this end, the rule ”d6 = 6d” is
applied. Using eq. (9) we obtain
i i
J (l + h — a) (l + h — a) v'Sv|o — J <5^ (/ + h — a)
v'Sv'da o v'da
o
i
/r\
<5^— (/ + h — a) v'da
as well as
11 t\ 11
JvSvdt = Mrfto ~ / i)8vdt J
to to vSvdt.
to
702 12. Variational principles in mechanics

The Hamilton-Ostrogradsky principle yields


to K o
SS h l9v' t) + i) (/,
-v +£)] (/,+ £)}
<7^ (/ l\ — .a) v' Svda-
(12.5.10)

Since variation 6v considered as function of time is arbitrary,


the integrand of the integral over t must be zero. However 6v
(a, t) as well as 8v (/, 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 = g-^[(l + h-a)v'] (12.5.11)

and, secondly, to the boundary condition at the end x = l


h [9vf (/, t) + v (/, t)\ = 0.
(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. l\ = 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,0) = f(a), v(a,0)=g(a), (12.5.13)

where / (a) and g (a) are continuous functions of a, such that /


(0) = 0 and
fl(0) = 0.
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, £), each of them being the
harmonic vibration referred to as the principal vibration
vk (a, t) = (a) (Ak coscukt + Bk sineukt).
(12.5.14)
Frequency ouk is not yet determined. Function <£>& (a)
termed the normal mode determines the amplitude
distribution along the chain and is the solution of the following
[(l + h-a)&k(a)}'+ ^$k(a)= 0, (fc = l , 2 , . . . ) (12.5.15)
12.5 Equations of motion of distributed systems 703

subject to boundary conditions.

$fc(0) = 0, *fc(i)-y*k(0 = 0, (A; = 1,2,...). (12.5.16)

Introducing the new independent variable a by means of the

equality

/ + h — a = 7-^-cr2
4
4
1d&k
da2 a da + = 0.
Its solution can be cast in the
form
(a) = ClkJ0 (a) + C2kN0 (a),
where Jo (a) and No (a) are Bessel functions of the first and
second kind, respectively. Noticing that
<4 (a) = [CikJo (4 + C2kK (41 ? = - [CxkJ'o (<r) + C2kN'Q (a)}
M
da ga
we can write the boundary conditions (16) in the form
CikJo (xk) 4- C2kN0 (xk) = 0,
C lk Jo (4) + 2XkJo (4) + C\2k K (4) + 2^0 (4) 0,
(12.5.1
7)
where

l + l\ l+h IX
Xk — 2cOk Xk Xk 1+g
(12.5.18)

The system of equations (17) has a nontrivial solution for


C\k,C2k if its determinant is equal to zero. Then we arrive at
the transcendental equation
Jo (xk) K(x*k) + -xlNo(x*k) - N0(xk) Jo (4) + 2XkJ° (4) = 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 N0{xk)Jo [XkJl (12.5.20)


l -\-l\
Vk (a, t) = (Ak cos ojkt + Bk sin u>kt),

so that
(a) = N0 (xk) J0 (^0^) - J0 (xfe) TVo ^xk^Jl-
The constants Ak and Bk are determined by means of the
initial conditions (13)

f(a) = YlVk (a> °) = 53 Ak®k (°) ’


k—1 k=1
OO ^ I--------XkBk
oo k (12.5.21)
9(a) = (°>°) = 2\jT+h'52 ®
k=1 k=1
These formulae are actually expansions of the prescribed
functions / (a)
and g (a) in terms of the normal modes $k (a).
The solution is based on the property of the generalised
orthogonality of
the system of functions <&k (a) expressed as follows
i
J (a) (a) da + (/) $TO (/) = 0 (k^m).
(12.5.22)
o
J (a) to differential
Indeed, turning (a) da + J [(/ + h - a)(15)
equation $'fe (a)]'
we $TO (a) da = 0.
9
After double integration by parts and using eq. (15) together
with the boundary condition (16) we obtain
L
^ K - a’l) J (a) $TO (a) da + /i$fc (/) 3>m = 0,
(/)
which yields relationship (22) for m ^ k. For m = k we introduce
the notation
L
= J ^ m(a)da + h^ m (/).
2 2
(12.5.23)
12.5 Equations of motion of distributed systems 705

Then, taking into account that due to eqs. (22) and (23)
i
J f(a)$m (a)da + f(l)$m (l)h
0
ooAk L
= ^2 <f>k$’mda + l1$k (l) $m (l)
k=1 — A N1v
AAm m
/
we obtain
(12.5.2
L 4)
Am = J f (a) (a) da + f (l) $m (l)
Nn
h
and by analogy
_0

’i
2 l + l\
B m %k, (12.5.25)
Nm 9 ■
J 9 (a) (a) da + hg (l) (l)
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 /x, however it is considerably simplified for the case in
which the end mass is absent, i.e. n = 0 or l\ — 0. Then
/ 1 2 l~a =
gcr ,
and a = 0 at the lower end of the chain. The solution of
l—
(a) = J0 (<t) Jo 2 uJk
a
9
since, if we kept Bessel function of the second kind, the
solution would be unbounded at a = 0. The frequency
equation obtained from the boundary condition (7) is

Jo (xk) = o, xk — (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
<ji = 1.2024 cj2 = 2.760 9 cj3 = 4.327 9
The normal modes take the form
$k (a) = Jo [xky 1 - jJ (k = 1,2,...). (12.5.27)
Equalities (22) and (23) yield

/ Jo ( \/ ~l) d<1={ Nm = Iff (xm)


Jo Xm 1 (k 7^ m)
0 (k = m) ’
where

N\ = 0.269/, N2 = 0.116/, N3 = 0.074/.


The solution is thus the following series
{a,l) = J
° (Xk\A “ j] (Ak cos~2]nt + Bk sin ~9~ \l f *
k=1 2V/
(12.5.28)
where
0
Ak lJo ( xk,/l - j ) da,
\j^ J 9 (a) (12.5.29)
Bb x
k\ll - y ] da-
XkNk Jo
The motion of the chain is a superposition of the principal
vibrations, each of them being a harmonic oscillation of
frequency Uk- The form of the k — th principal vibration is
given by function (a). According to expression (27), this
function has, besides a = 0, another k — 1 roots, given by the
equalities
(to = 1,2,... —1).
(12.5.30)
These determine the nodal points of the k—th vibration mode.
For example, the second mode has a single node
= l (\ - ^ sa 0.81/,
the third mode has two
<43)nodes
= 0.923/, 43) = 0.595/
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 O
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 uj. The other end of
the rod x = l 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' = ii (u + R + a) + i2v + i3w.
In Fig. 12.5 point M' shows the actual position of point M of
the undeformed rod. In the latter equation a = OM and u, v, w
are 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

v! = -1 +t/) . (12.5.31)

v = vr + u x r' = uii + vi2 + wi3 + UJ\3 xr',

where vr denotes its relative velocity. Then we obtain


vx—u — ujv, vy = v + u (u + R + a), vz = ic,
708 12. Variational principles in mechanics

and the kinetic energy of the rod is given by i l

T = ip J (yl + Vy+ v2) da = j da \uj2v2 + v2 + 2uv (R + a) +


o

o
Here only terms of second order of smallness are kept
(please, notice that u is a second order quantity whereas v
and w are first order quantities). Since
i
1 \ j /H' a I au'da
J (R + a) uda R+ -a J au|0

L
= ~\i R + — l j l — (r + —a ) a (^vf +w'^J da

we can cancel out the nonessential constant term in the


expression for T,
to obtain
i
T = J da {lu2v2 + v2 + 2luv (R +a)—
UJ R(l — a) + a- 2(l) z—
2

(vf + wf ^ + w21.

(12.5.32)
l
n =^J (EIZV"2 + ElyUi"2^ da, (12.5.33)
0
and Hamilton’s action takes the
form
S = J dt 11J i da [v2 + w2 + uj2v2 + 2ujv (7? + a) —

(12.5.34)
to o
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

The differential equations of vibration and the boundary


conditions at x = l are obtained from the condition 6S — 0

8S = p J dt J da \v8v + w8w + uj2v8v + u (R + a) 8v-


10 o
2 R + 1 + af , f ff EIZ EL
uj2 (l — a) (v 8v + w'8w' -v"8v" ~~ w"8wf 0.
(12.5.3
6)
Integrating by parts and taking into account the equalities
8v (a, to) = 0, 8w (a, to) — 0, 8v (a, ti) = 0, the (a, ti)
= 0, th/ (a, to) = 0, 6w' (a, to) = 0, th/ (a, ti) = 0,
trn/ (a, ti) = 0,
we can recast eq. (36) in the
form
ti i , ,7 N 2 R + / + a f (l EL 1vIV \ 6v
— a)-------------v
8S = pt oJ dt
0 J da < — v + u v +
2

ti l ,
2 R T / T a f EL -wIV > 8w
J2 (l — a) w
+ P / dt J dal —w +
to 0 ^
11 m
J
+ EL dt [v (l, t) 6v (l, t) — v" (l, t) 8v' (l, 0]
to
tl

+ Ely J dt [w,n (/, t) 8w (/, t) — w" (/, t) 8wf (/, t)] = 0.

Finally, we arrive at the following partial differential equations


and boundary conditions: for vibration in plane Oxy
EIZ 1 v n N 2 R v+ / + a f
v -|----- v — u z
v — (l-a) o 0, > (12.5.37)
ujz
v (0, t) = 0, v' (0, t) = 0, v" (/, t) =
0, vm (/, t) = 0 and for vibration in
plane Oxz
EL , . 2i£ + / + a .
w H----v — ujz (Z — a)- - -------w = 0,
(12.5.38)
P
u; (0, t) = 0, w' (0, t) = 0, u/' (/, t) = 0, w (/, t) = 0.
,n

Integration of these equations was carried out in Sec. 9.10.


710 12. Variational principles in mechanics

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
a a

11 = 7 J yda = 7 J y\J\ + y'2dx.

Here 7, ±a and da — \/l + y'2 dx 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= yy/lT
which does not contain the independent variable x. The
’’equation of motion”, which is Euler’s equation for the
variational problem under consideration, has the energy
integral (7.2.5). In our case it is given by
y
= -yo,
\/l + 2/'2
where the constant y0 denotes the value of y at the lowest
point of the equilibrium form (where yf — 0). Hence,
12.5 Equations of motion of distributed systems 711

and differentiating yields


„Vn
y —2 = °-
Vo
The solution, which is even and equal to yo at
x = 0, is
V = 2/0 cosh (12.5.39)
—.
yo
Then we have
X X
yf = sinh —, dcr = cosh —dx, (12.5.40)
Vo
and the length of the chain is
a
7i^xi^ ..a l../2al (12.5.41)
/■ —dx = 2yo sum — or ---------= sum — - -----
l cosh
yo yo tyo \ l 2 y0
The unknown value —should be found from this transcendental
equation 22/o
for a given parameter y.
In what follows we will need the formulae

dx 1 (12.5.42)
cos ce cosh — sin a = tanh ■
yo yo

where a denotes the angle of the tangent to axis Ox. From them we obtain
dx da 11
da y0 cosh (12.5.43)
—, da P yo cosh —
yo
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 are its projections on the
tangent,
u = /xt T ism T /3b, r = r0 + 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 |dr| = da, which implies that


v dp 2 LL dlS 2
+ +
—TT”
P 0(7

After simple manipulations we have

-~ir + \\(t + ir)2+(--7r)2+(f)2 (12.5.45)


p da 2 \p da J \p da J \da)
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 da da
(12.5.46)
p da da
Keeping the second order terms we can cast the constraint
equation as follows
dp J_ d 2p
P + drf (12.5.47)
da 2
We proceedpto construct the equation for the potential
energy. By virtue of eq. (44) the vertical coordinate of the chain
point is
rj = y + psin a + is cos a,
where y denotes the coordinate of this point on the chain line
at the equilibrium. Then, taking into account the inextensibility
of the chain we have
n 7 J rjda = 7 J yda + 7 J (/x sin a + is cos a) da

= n0 +7 J (/i sin a-\- is cos a) pda,


-a0
with n0 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 is by eq. (47) yields
CKO
' . dp
n-n0 p sm a + -g- pda ~b
cos a
— CKO
2 dp 2'
+ da da.
— CKq
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
c*o
[ ( . dy \ f( x dy \ x
I /isma + Tjj— cos a 1 pda = [ ytanh- - - -\-—yo l ) cosh—ax
yo dx yo
— C*0
a0
[—
yo J dx //cosh — \ dx —
0,
yo'
because y = 0 at x = ±a. Thus,

n - n0 = I7 J cos a -so +(£ da, (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

— CKO
While deriving this result we used relationship (46) and
neglected the fourth-order terms.
Now, entering the non-dimensional variables
X
T= y
o
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
Co
L b I fi2 cosh £ + fir cosh3 £ - ( (12.5.50)
~ Co

Co
P
cosh £ + y! sinh £)2 d£ + ^7 J cosh £

Co
Here eqs. (42) and (43) are used and £0 denotes the value of
£ at x — a. The required functions y and (3 satisfy the
boundary conditions
at £ = ±£0 y = 0, y! — 0, ft = 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. 8S = 0, and
integrating by parts in order to remove the derivatives of
variations 8fi and 8(3 by means of the boundary conditions for
the variables, we obtain the differential equations for small
vibration of the chain
ji — ji" cosh2 £ — 3jif cosh £ sinh £ + 2// sinh £ + 5fi" cosh £ +
4//" sinh £ + fxIV cosh £ = 0,
/?cosh£ — /?" = 0, (12.5.53)
Taking
fi = M (£) sin (At + a), /? = B (£) sin (cut + 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
MIV cosh £ + 4M'n sinh £ + (5 + A2 cosh £) M"
cosh £+
(12.5.54)
(2 + 3A2 cosh £) M' sinh £ - A2M
= 0, ►
M(±£o) = 0, M'(±£ o) = 0, ^
and for vibration perpendicular to the plane of the (12.5.55)
chain
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 equations for small vibration of one-dimensional
bodies about the equilibrium position in the potential field. The
equations obtained are partial differential equations for
unknown function v (x, t), with x and t denoting respectively
the coordinate of the point in the equilibrium configuration and
time. The solution had two steps. First, we introduced a family
of partial solutions
vk (x, t) = Vk (x) (Ck cos u>kt + Dk sin ojkt) (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 uJk 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 boundary conditions provides us with such values of uo\
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 was 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 approximations to determine the frequencies and
modes of vibration. In practical applications it is important to
know a few first frequencies and modes, predominantly, the
fundamental one. In the framework of this approach the
distributed system is replaced by a system with a finite
degrees of freedom 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
nn ^nn
r
=5££ dskQkqs, n ^ ^ ^ CskQsQk-
(12.6.2)
Z
s=1 k=1 Z
s=l k=1

It is assumed that, under specially chosen initial conditions,

qs = Cs sin (cut + a) (s = 1,... , n),


(12.6.3)

where all of the generalised coordinates have the same


phase uot-\- a. Such vibrations are referred to as principal
vibrations. The problem is to obtain the system of algebraic
n
£* (L) = X] (a*k<ik + Cskqk) = o (s = l,..., n). (12.6.4)
k=l
716 12. Variational principles in mechanics

Substituting expression (3) for the generalised coordinates


yields the required relationships
n
^2Ck(-Lo2ask + csk) = 0 (s = 1,... ,n).
(12.6.5)
k =1

However, the idea is to avoid constructing the differential


equations and to find the required values directly from the
Hamilton-Ostrogradsky principle 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
Hence,
Sqs = 8CS sin (ut + a) + (Cs8a + Cst8uo) cos (out + a) (s =
1,... , n)
(12.6.
6)
The Hamilton-Ostrogradsky principle should be written in the
form
ti

J 6Ldt = 0,
(12.6.7)
to

which is equivalent to the requirement that


ti
6S = 0, S = j Ldt
(12.6.8)
to

for fixed d t\. Provided that and t\ depend on the varied


an

quantities, the above statements are no longer equivalent.


Then, taking into account that the integration limits are varied,
we have
ti
SS = J SLdt + (L) t=ti Sti - (L)t=t0 Sto,
(12.6.9)
12.6 Approximate determination of natural frequencies and normal forms
717
Let us assume that t\ and to differ by one period, i.e.
2tt 27r
t\ = to H- - - -, 6ti = Sto (12.6.11)
uo^Suo.
--------------- uo
By virtue of the relationship
tl n t!
j 8Ldt = - J 2 j s s ( L ) 8qsdt +
to S= 1to

0, (12.6.12)
E
S—
1
we can adopt that
6qs (h) - 8qs (t0) = 0 (s=l,...,n),
(12.6.13)
dL
due to the periodicity of 7— rather than consider that each of
these vari-
. . dqs
8S+(L)t=to-^8w = t). (12.6.14)

Returning to eq. (6) we have


6qs (ti) = 6CS sin (uoto + a) + Cs6a cos (uoto + a) + CstiSuo cos
(uoto + a), 6qs (to) = bCs sin (uoto + a) + Cs6a cos (uoto + a) +
CstoSuo cos (uoto + a).
For condition (13) to be satisfied for to~ti ^Owe should adopt
that along the neighbouring paths
uot0 + a =
(12.6.15)
Substituting the generalised coordinates (3) into the
expression for L 2yields 2
L = uo T cos (uoto + a) — U sin2 (uoto + a). (12.6.16)

Here T and U are quadratic forms obtained from T and n by


replacing qs
and qs by Cs
^nn ^nn
(12.6.17)
r=
2EE »
a tC
^' u =-Y,Y. ° ° -
c kC Ck
s—1 k— 1 s=1 k=1
718 12. Variational principles in mechanics

They are referred to as the amplitudes of the kinetic and


potential energies. Then we obtain
Cl
S = J Ldt = ^ (w r - u)
2 (12.6.18)
to

and
ss = 7T (r + T) 6w + - (w2^r - su) (12.6.19)
UJ* J
UJ
(Vt=t0

we can write the Hamilton-Ostrogradsky principle (14) in the


form
r - L ) 8w+- (w2<5r - SU) = 0. (12.6.20)
UJ* J
UJ
Let us make use of the energy integral
cj2rcos2 (ujto + a) + U sin2 (ujto + a) = const,
where the constant on the right hand side is determined by the
left hand
side at t = t$. Then we have
or
uj2T = [/, (12.6.21)

and it is possible to be represent relationships (20) as follows


lj26T -6U = 0. (12.6.22)
Thus, if we introduce the quantity
R=-S = u?T- U, (12.6.23)
7r
its variation S'R is zero for the fixed cj, i.e.

gR
=“2<r -su=£ ~ H)6C-='“• '
(12 6 24)

v 7
s=l
Since variations SCs are arbitrary we obtain the system of
equations (5) by means of eq. (17). The expression for R does
not contain the initial
ar dU k ( 2 dT dU
^ sJ ) Qsr = }^
dCs ' " dC dfir ' d^r
\

J <Vr-
(12.6.27)
Since 6fir are arbitrary we come to the homogeneous system
2 dr du n . ., of equations
u?--------- =0 (r = 1,...1,k) (21.6.28)
OjJLr OfJ,r
which are linear in fir.
The requirement of a zero determinant leads to the
frequency equation
which is an equation of the k — th power of uj1 and has, in
720 12. Variational principles in mechanics

different values which are all positive. For each value of


there
is a system of values (i^,... , determined up to an arbitrary
factor. Then, by virtue of eq. (26) we can determine the
coefficients d(up to an arbitrary factor) corresponding to the
normal mode with frequency ur.
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 q sr 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 uj\ and the first normal mode. If
one takes k = 4 then one expects satisfactory accurate
information about uj\ 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=^J p(x)F(v,v')dx, U = \J &(v,v',v")dx, (12.6.29)
0 0

where p (x) is the mass per unit length and F and are
quadratic forms of their arguments. Assuming

v (x, t) — V (x) sin (<jjt + a),

(12.6.30)

we obtain expression (16) for L where


i i
r = ±Jp(x) (V, V') dx,
F
(V, V', V") dx. (12.6.31)
\j p{x)
9F„, <>F ,, , (12.6.32)
ST = 0 Wsv+wsv Ux
1

' , OF
d dF , 1 dF 6Vdx+
c rJ
^Pgyl^y
^dV-Tx^W' 0
1 Y d<f> d d<P \ 6V+
„ d<& „ .1
) dV"6V\
2 \dV ~ dxdV" j

(12.6.33)

Then we
obtain
dF d dF\ d$ d d§ d2 d<f>
' dV dxPdV'
U P SVdx
dV dxdV’ dx2dV"
gR
= \)
1 + o dF d4> d dd> dV"
2 uj pdV' dV' dxdV" + 8V - 8V = 0 (12.6.34)

and the problem is reduced to the homogeneous differential equation


dF d dF\ <94> d dd> d2 dd>
0 (12.6.35)
dV+dxdV' dx2d V"
1
" sv - ^W'
subject to two sets of boundary conditions at x = 0 and x = l.
The first set is: either
dF d$ d d$ n
u> p- — + —^7 = 0 (12.6.36
dV' dV' dxdV" )
or
V=
(12.6.37
whereas the second set is: either )
0:
<94>
0
dV'
or (12.6.38
)
V' = 0.
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 (12.6.40)
k
V = ^2 dvFr Y) ,
722 12. Variational principles in mechanics

where cpr (x) are some functions satisfying the geometrical


boundary conditions. Substituting this expression into
formulae (31) and estimating the integrals we obtain T and U
as quadratic form of parameters /i1?...
Then we construct equations (28) which yield k frequencies
uj\, ... , uj\ and k sets of parameters ,... , /xj^ 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 T and
U with further construction of equations (28) we turn our
attention to equality (34) and replace variation 8V as follows
k
SV = Y,Vr(x)6/ir. (12.6.41)
r=1

Then we obtain equations (28) by equating the coefficients of


i
2dV dU 1 f r 2 / &F d OF \ d<$> d d$
^ dJTr ~ d^ ~ 2 J f \PdV ~ dxPdV'J dV + dx dV'
d2 <9$ dF d<!> d d$ ^
ipr (x) dx + - u) p Fr (x)
dx2 dV" dV' dV' dx dV")
+ ~QyJJ(Pr(X)
0 = (12.6.42)

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 dF d<t d d$ d$ x 1 l
2 u) pdV'> Fr (x) ~ dV" Fr ( ) (12.6.43)
dx dV" -0
W'
vanish. Some of these vanish due to the geometrical
boundary conditions since functions ipr (x) are subject to these
conditions. The remaining terms vanish provided that ip r (x)
are subject to force boundary conditions. Then equation (42)
takes the form
1 d dF\ d<S> d d<f> d?
2 diPdV!) ~ dV + dxdV' <94> ipr (x) dx = 0
dx2 dV"
(r = l,... ,fe), (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 equation
^(y,y/,y//,...) = o,
the sought-for function V is replaced by its approximation
(40), functions (pr (x) satisfying all the boundary conditions.
Then by means of the following relationships

o
whose number coincides with the number of the taken
functions ipr (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

(12.7.1)

hence, by virtue of eq. (6.23) we have

(12.7.2)

Let us search V in the form of the polynomial


V
- Miy +/i2p- +^3]3 (12.7.3)
where functions
(*=1,2,3)
724 12. Variational principles in mechanics

satisfy the geometrical boundary condition V (0) = 0. After


elementary, but bulky, calculation we obtain the following
quadratic form of parameters fik
R {"■ (t + ^]+/x|(-+^j+/X3^-+^) +
+
2^ 1M2 ( 4 + j
H +
2^3 ( g
^ +
f 5 +K ) ""

9
lid2 Mi ( 2
+ +
M2 5+rJ+^u+5"> +
2M1M2 ( q + + 2/x2M3 ^Yo + 2™) + ^M3Mi ^4 + ^ | > (12.7.4)

where x denotes the ratio of the end mass and mass of the
chain (designated by fi in Sec. 12.5). Equations (6.28) are
easy to construct now. Equating their determinant to zero
yields cubic equations for a;2, all roots being positive. Then we
obtain three modes Vr (x), among them V\ (x) and V2 (x) being
sufficiently accurate. As the calculation is very cumbersome in
the case x 7^ 0 we restrict our attention to the first
approximation by keeping only the term in eq. (3)
V (x) = Miy-
This corresponds to replacing the chain by a rigid rod. Then

R= l +x 9
Ilo
2

and the approximate square of the frequency is


given by
o 3 g 1 + 2x (12.7.5)
uz =--------------.
2/1 + 3*:
Clearly, factor /i1 remains undetermined. If x = 0
we have

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 Vl(a) = y,
$1 ia) — Jo Xi 1 1

so that at the midspan

+++H-4’ +)=0-5-
a/l $1 (a) Vi (a)
0 0 0
0.25 0.178 0.177
0.50 0.398 0.403
0.75 0.670 0.678
1.0 1.00 1.000
0 0

Good agreement is observed. In order to obtain the second


frequency
and the second mode to satisfactory accuracy it is necessary
to keep four
726 12. Variational principles in mechanics

12.7.2 Vibration of a rotating elastic rod


The comparison is restricted to the bending vibrations in the
Oxy plane. Using the expression for 8S of Sec. 12.5 we obtain

(i ^ + l ~ aT// f \
6'R = p \ (uo2 +\2)V{a) + uo2 (l~a)-------5----V (a)
o
FT
~yVIV (a) l da + EIZ [V" (l) 6V (l) - V” (l) 8V’ (/)] = 0, (12.7.7)

where A2 denotes the square of the required eigenfrequency.


Taking
a
2
2Q 3 r
Q Q Q 2
^ = Mi p + -
^~ /3 bV = 2—8+ 3—

(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 8Hi and 8f±2 to zero we arrive at the
l -i / | o
h (ca 2
+ A2) V + ca2(/-a)gg+2 +v
/
1 z
0 a
¥+
EIZ
l P V" (l ) - jV" (l ) = 0,
(12.7.9)
n 2 ^
/ da (<-jj + A ) V + ljG-a)------j^
2 2 + ^ + a
T// -I / | Q
1 1
Z3
EIZ Vm (/) - jV" (/)
P 0,

since Viv (a) = 0 under the adopted approximation.


To begin with, we take J1 = 0, i.e. we consider a non-rotating
homogeneous 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
(ib _ 4)+ ^ Qz - 6^ = o,
Mi Uz - 6J + n2 [jz - 12) = 0, (12.7.10)

where

z=
12.7 Examples of approximate calculation of natural frequencies and forms
727
Having constructed the frequency equations we find the
smaller root
I El
A = 3.53 J-jZ,
(12.7.11)

J_ 1R ,.4 3R
15 + 37 -z — 12 ^ *35 + 10 l
127 12
0, , pi
4
-z-6-ip 2 (12.7.12)
3_R * =
M W

6 +
io 7
which enables estimation of the influence of angular velocity
on the fundamental 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
/ (/?2 (p4 \
n = mgl (1 - cos ip) = mgl ( ^— — + ... J
we keep the above terms and introduce the non-dimensional
independent variable which is denoted below as t. Then, the
kinetic potential, up to the constant factor, is as follows

L
=^2-^2 + Y4- (12-7-13)
The motion of the pendulum under the initial conditions
at t = 0 ip — ip0, (p = 0.

(12.7.14)
is required.
Within the linear approximation the solution has the form
ip = ip0 cos t.
(12.7.15)
If the last term in eq. (13) is retained, the frequency of vibration
Using the Hamilton-Ostrogradsky principle we can find A and
a. We have
6ip = 6a (cos At — cos 3At) — tSX [(ip0 + a) sin At — 3a sin
728 12. Variational principles in mechanics

so that 6(p is zero at t = to = 0 and t = t\ = — = 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
SS - (L)t=T 6T = 6S + (L)t=0 6X = 0. (12.7.17)
A
Then we have
Cp — — A ((p0 + a) sin At + 3Aa sin 3 At (12.7.18)
and thus
/r\ 7T 7rSX {1 9 1 4^
(^=0^ = --^^--^.

Then, inserting eqs. (16) and (18) into eq. (13), we find
i
S — J* Ldt = — [A + 2(foot. + 10a2) —
o

- (ip 2 0 + 2(p 0 a + 2a
2
) +— oa (12.7.19)

It will be shown later that the parameter a has the order of


y?g. With this in view we have omitted terms higher than
second order in a and products of a2 and (p0 in the above
expression. Furthermore we have
6S+(L)t=0^6\ =

(<Po + + 10a2) + ^2 (-<Po + 2(^oa + 2q;2 - 6


7T A (2(^o + 20a) — — \ 2(p0 + 4a —
+
4 A

Cancelling out the terms proportional to a2, two equations


5a =for
0. a
and A are obtained
1( 5
(pi + 2(p 0 a + -^2 ( —Vo + 2 ( Po a ~ 0,
43^0
ip0 + 10a - ^ yPo + 2a - ) = 0.

Keeping only the lowest order terms in (p0 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
, -i Vo\f , cos ( 1 - ^|) f - cos3 (l - )(
¥) = *>„ cos I 1 - I , + J92
(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
L — ^ ^ PsQs H (<7l 5 • • • 5 Qn 5 Pi 5 • • • 5 Pn 5 ^) •
(12.8.1)
s=l

Then
t
^2psqs - H (qi,... ,qn,Pi, ■ ■ ■ ,Pn',t)
/ dt. (12.8.2)
'S= 1 L

and the variation of this integral is


given by
t
6S (ps<% + qsdps ~ *r!~~hq
= J s - ^~dps ) dt.
dqs dps
(12.8.3)
to

By using the canonical equations of


motion
. dH . dH . , , (12.8.4)
Qs o 5 Ps - - r\ l? * * * 5
W')-5
we transform equality (3) to the
form
t" _n t pn n

/y,3 Sap Sq). + Psbq.s) dt = I d.y p Sq = y (p 6q.s ~ /


s
( s
s
s s s s

(12.8.5)
Here as and (3S denote the initial values of the generalised
coordinates and momenta
at t — to qs Ps 1,... , ti) .
(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 canonical transformation of the type
730 12. Variational principles in mechanics

of the actual and initial values of coordinates, time t and the


fixed time instant to
S = S (qi,... ,qn,ai,... ,an;t,t0), (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
aq
ir=Ps, 7T- = -Ps (* = !»••• >«)•
(12.8.8)
UQS 00's
After solving these equations for the coordinates q s and
momenta ps
which implies a non-trivial Gessian (12.8.9)

d2S d2S
dqidai ’’' dq\dan
G= .............................. ,
d2S d2S
dqndai ''' dqndan (12.8.10)
we come to the equalities
Qs (MOj^l, • • • 5 5 fil 5 • • • 5
qs
fin) ’\
Ps Ps ifl ^0) ^1? • • • 5 ^77.5 fill • •
• 5 fin) 5 J
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
d dS ^dS . ^ . rr dS .
S 2^TfcTfc 2^Psqs ~H = Z^ H.
dt dt + a=1 dq
* i=i= dq
°
12.8 Hamilton’s principal function
731

The sums on the left and right hand sides cancel out and we
obtain

£+«(•..............<-•“>
Hence, the principal function satisfies the Jacobi equations and
represents the compete integral (10.13.27) for which the
constants aq,... , an are equal to the initial values of the
generalised coordinates q\,... ,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 (3k 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 complete integral of Jacobi’s equation
V = 'S(q1,... ,c?n,7i,... ,7„;0+7n+i>
(12.8.12)
v= (qi,... ,qn,7i,- ■■ iln’i) ~ ® («i> - •• ,<*n,7i,-- - ,7„;<o) •
(12.8.1
3)
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
dV_ = d'il (qi,... ,qn, 7i,--- ,7ni0 _ d(^au... ,ct»,71,... ,7 n;t0)
dlk ®lk dlk
= 0, (fc = 1,... , n).
(12.8.14)
The generalised coordinates q\,... , qn are obtained from this
system in terms of the initial values aq,... , an, time and
constants 7i, • • • ,7n- Excluding 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 motion of a particle in a vacuum. Using expressions
(10.12.9) for the integral of Cauchy’s problem we obtain
t
s=
\ / ( * y2 ~ 29V) dt
2+
p= X- Xo
t -to ’
1f x —
h=
2U-

(12.8.18)
+ \g{y + yo) + \ (jzjf) + \a2 (* - to)2
-1 0 .. . 0
0 -1 .. . 0 = (-l)n, (12.8.22)
0 0 .. . -1

where A (£, to) is determined by eq. (3.20) and reduces to


zero in the conjugate 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 + At). The difference
q*s (*) - 9. (t) = 6qs
(12.9.1)
is the synchronous variation introduced above. Taking into
account only values of first order of smallness, we have
q* (t + At) = q* (t) + q* (t) At = qs (t) + 8qs + qs (t) At, (12.9.2)

so that
qs (t + At) - qs (t) = Aqs (t) = Sqs + qs (t) At (s = 1,... , n).
(12.9.3)
This equality defines the asynchronous variation denoted by
In particular,
A qs = 6qs + qsAt = (6q3)m + qsAt. (12.9.5)
The value At in eq. (3) is an arbitrary differentiable
infinitesimal function of time. Hence,
(A/)# = («/)# + f At + / (A^)# ,

(12.9.6)
so that
(Aqs)m = (Sqs)9 + qaAt + qs
(At)9 = Aqs + qs (At)9 . (12.9.7)
Therefore, the operations A and d are not commutative, in
contrast to 6 and d.
Applying formula (4) to the integral
t
j Fdt,
0
we arrive at the equality
t t
A / Fdt = 8I Fdt “I- FAt. (12.9.8)
12.10 The Lagrange principle of stationary action 735
On the other hand,
t t t

J AFdt = J (6F + FAtj dt = J [6F - F (Af)*] dt + FAf . (12.9.9)


00
Removing the
expression t
6 j Fdt = J SFdt,
o o
from these equalities yields the following
relationship
t t
(12.9.10)
A J Fdt = J [AF + F (At)*] dt + (FAi)t=lo •
0 0
Since
t\ t\ to

J Fdt = J Fdt - J Fdt,


t0 0 0
we immediately obtain
ti ti
AI Fdt = J [AF + F (At)*] dt. (12.9.11)

12.10 The Lagrange principle of stationary action


Let us consider a material system subject to holonomic
stationary constraints. The active forces are assumed to be
potential forces so that the total mechanical energy retains a
constant value throughout the motion
T + Il = h. (12.10.1)

It is adopted that this relationship remains valid for all


neighbouring paths passing through the two fixed positions
qi°^ and 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
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

Let x\ denote the position of the particle at instant t\ when the


particle moves along the true path. It is not possible to arrive
at the same position x\ 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 of the forthcoming analysis we notice the
relationship which is obtained by varying equality (1)
accounting for formula (9.4)
a (t + n) - 6 (t + n) + (t + n)# At - s (t + n) = Ah = sh = o.
(12.10.
2)
Let us turn now d nto Lagrange’s fundamental equation
— ^p.Ms = 6L = 6T-6U = 2 ST.
S= 1
Replacing here 8qs by means of eq. (9.3) and noticing that for
the stationary constraints
n
^Psils = 2 T,
s=1
we obtain, due to eq. (9.4), that
jn i
— y^psAqs - 28T + —2TAt = 2ST + 2TAt + 2T (At)* at at
s= 1
= 2A* + 2T(A*)V
(12.10.3)
Integrating now both sides of the latter equation from to to
ti, recalling formula (9.11) and noticing that
Aqs(t0) = 0, Aqs(t1)=0 (s = 1, (12.10.4)
as the neighbouring paths intersect the true path at time
instants to and £1, we arrive at the equality
12.10 The Lagrange principle of stationary action 737

The value

A=
2 Tdt (12.10.6)
to
is referred to as Lagrange ’s action. For a single
particle we have

mvds (12.10.7)

where the integration limits to and t\ 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
(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 suggested 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.
(12.10.9)

where
F = 2T + A (T + n — h) (2.10.10)
738 12. Variational principles in mechanics

and A denotes Lagrange’s multiplier.


By virtue of eqs. (9.1) and (9.4) we have

A\I> = J [AF + F (At)*] dt = J [$F + FAt + F (At)*] dt


to to
ti ti
= J [6F + F (At)'] dt = J SFdt+ FAt\o.

The integral is transformed by means of formulae (1.4),


(1.8) and (1.9). Applying also formula (9.3) we obtain

V 2-Sq, + FAt
ll
n
r

A9 = - ^2 £s(F)6qsdt + ti8qs .
to s=1 to
n
= -/Z)^(^)^+
L 8= 1
to

The latter term vanishes due to condition (4).


Hence,

A9 = - lf'££s(F)6ql dt ~b rz&r.
ti
S=1
= 0. (12.10.11)
to

As (At)0 and (At)x are independent and we seek the stationary


value of functional 4/ containing Lagrange’s multiplier A, the
factors of (At)0 and (At)1 are equal to zero, as well as the factor
of each 6qs in the integrand. Then we obtain, first, n
Lagrange’s equations for F
„ d dF dF ^ .
£s (F) = —-------x— = 0 (s = 1,... , n) (12.10.12)
dt dq dq s s

and, second, the conditions at the end-points

r-±W*) -0 ’ =° (i2.io.i3)
^ s—
1 /t t=t 1

Let us recall that a consequence of Lagrange’s equations is


eq. (7.2.4), which takes the following form
d/ F ”dF.\ dF {/rT1 „ f, „
-E^-J =^ = A(T + n-l,) = 0.
dt
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 = hence it is held constant for any t

(12.10.14)

Inserting the expression for F into this equation we find


2T + A (T + n - ft) - (4T + 2AT) = -2T (1 + A) = 0.
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 asynchronous variations if we exclude time from the
explicit expression of the principle of stationary action. This
possibility is given by the energy integral (10.1) which can be
written in thenform
n
2 T(dtf = ^2^2ask(qi,--- ,qn)dqsdqk = 2 (h - II) (dt)2 . (12.11.1)
s= 1 k=1
This enables dt to be expressed in terms of differentials of the coordinates

(12.11.2)

It is substituted into the expression for Lagrange’s action to give

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 qi, as the independent variable.
Then A is set as follows
(i) (1)
f da = a^
J dcr da = j <$>da =
(0 ) (0) (0)

-a<°> (12.11.10)
12.11 Jacobi’s principle of stationary action 741

where we have introduced the quadratic form

$ - bsk dqk
s
1. (12.11.11)
dq da
da
The ”arc” cr, taken as the independent variable, represents
Lagrange’s action 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 i?* is the
extremal of functional (10). We can immediately write down
the differential equations by using the above analogy and re-
placing L and t by <I> and cr, respectively
d d$ da dqs 0 (s = 1,... , n). (12.11.12)
gdqs da

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
equations1 have the form, see eq. (B.8.7)
d qa d'f dip
da2 __=0 («=!,... ,»), (12.11.13)

where Christoffel’s symbols are constructed for manifold R


with quadratic form (8).
The independent variable a does not appear explicitly in the
expression for <I> and this allows us to adopt, without loss of
generality, that a — 0 and a = a\ in the initial and actual
positions of the system, respectively. In addition to the given
constant /i, the general integral of system (12) has 2 n
constants csl
q = qs ,c2„]h) (s = l,... ,n). (12.11.14)

To determine the 2n + 1 constants c\,... , C2n; h we have 2n


boundary conditions
(q ) = <f(0,ci,... ,c2n\h),
s {0)
(<f)(1) = q {auci,... ,c2n\h)
s
(12.11.1
5)
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 integration constants of eq. (12).
The equation for time, due to eqs. (2) and (8), is as follows

f da (12.11.16)
t = to +
j
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 constraint can be related to the motion of
representative point on the Rieman- nian manifold Rn. Its arc
ds2 = 2 Tdt2 = askdqsdqk. (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
2
da = bskdqsdqk = 2 (h — n) ds2 = 2 (h — n) askdqsdqk. (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
bsk — 2 (h n) 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
|&sfc| = [2 (h - II)]” \ask\, Bsk = [2 (h- n)]”-1 Ask
so that the relationship between the contravariant components
of the metric tensors in R* and Rn is as follows

(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
bskasak = ask^J2 (h - Yi)asy/2 (h — U)ak = a,sk@'S@/k
= bskasak ^sk O'k askasdk, (12.12.5)
^2(h-n) y/2(h-TT)
which allows us to relate vector a in R* to vector a in Rn due to
A/2 (h - II)as, as = the rule
a
* (12.12.6) y/Hh-ny

the factor y/2 (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 bsk vector
Q'sk® ■y/2(h^TV)as = ak = as
2 (h - n) v v ’ y/2 (h -
n)
The scalar product a • b is invariant since
a•b=a
s
bs = asbs = a • b,
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 ther« metrics of manifolds R* and Rn is stated by the formula
i 7 a dbPl
1 db
[ft7;o1
( P<* Ob. (12.12.7)
= 2 (a? -
dqP dq
— 2 (/i II) [/3, y, Clj (c^/3aiII7 T T QL/3'yHQ:) 5
where the sign ^ designates the value in the metric of the
kinematic element and IIQ, is the derivative of II with respect to
qa. The relationship between the symbols of the second kind is

{“} = [A2;«! = {“}-Sfib) + <n-> + »^n“)•


(12.12.8)
Let us consider a certain trajectory of the material system
which is a geodesic of manifold R* and let r denote the unit
vector of the tangent to the geodesic, then according to eqs.
(2) and (6)
dqa dqa (12.12.9)
T= da yj2 (h - n) ds y/2 (h - n) ‘
744 12. Variational principles in mechanics

The differential equations (11.12) are equations of the


parallel translation of vector r along the trajectory and express
the fact that the geodesic curvature becomes zero. In the
covariant notation they have the form
dr13 n
6a/3-^-+ [/?,7;a] r^r7 = 0.
(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
is zero. For the forthcoming analysis it is necessary to
introduce the unit normal vectors c which are orthogonal to r
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
dcP a
baf3— + [/3,'y;a}cl3T~t = 0, (12.12.11)

which is analogous to eq. (10). One particular solution is


,3 ,k
d 3 k dc k j dc ^
(12.12.12)
da C c = ^■
c+c

For this reason, while considering the system of (n — 1)


independent
. \ TL \ TL —
1

particular solutions c ,... , ca of differential equations (11) for


a

(c)o • (C)o= ( boc0 CQC^ r o (k^j), l 1 (k = (12.12.13)


j).
1 TL — 1

The system of vectors r, c,... , c form an orthogonal basis of n


unit vectors belonging to manifold i2* . These unit vectors
move along the trajectory and do not change their directions,
S=°' ^ = ° (‘ = 1. ■ •»-!)■
(12.12.14)
They are determined up to a rotation about r. If vector p can
be represented on this manifold in the form
1 n—
1 (12.12.15)
P = (0) + <2(1) c + . . . +
a T
C,
12.12 Metric of the element of action and metric of the kinematic element
745
then, its derivative with respect to a is given by
— a(0)T + c + ... + a'(n_i) c ,
(12.12.16)
where a prime denotes a derivative with respect to a. In this
regard, this orthogonal basis replaces the axes of fixed
y/2(h- n)aa/3 + [P,T, alfdf1
ds y2 (h - n)
1 &Z-7 (12.12.1
2 (h - n)
^_ + v,ra]^-2{h_u) y.ds 1 <m „
rv ; 7)
= 0.
Here we used the following relationships
dU
ap^yT^T 1 1
= 1, a^n^f^f7 = fallpfP = fa gradn • f = fa —,
(12.12.18)
because are the covariant components of vector gradn.
Noticing that
ap^cf3 = t • c = 0 (12.12.19)
and carrying out the analogous transformation of eq. (11) we
obtain

aap^ + WPM&t'' - = 0.

(12.12.20)
dr 1 1TT ~dn dc_=Tgrad n • c
ds grad n — t Ts 2{h-n)' (12.12.21)
2 (h - n) —— ds
Recalling Frenet’s first formula and taking into account that 2
(h — n) and grad n are respectively equal to the square of the
velocity of the representative 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
= ^2 (Q - Q • tt) = ^2 Q • fin,
(12.12.22)
where n and denote the first normal to the trajectory and the
first curvature, respectively. Clearly, fc(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 a has the direction of f 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 L n—1
r, c,... , c at point M of the trajectory. Let us consider two
infinitesimally close points N and N' in the vicinity of M. Let N
and N' lie in the hyperplanes which are normal to the trajectory
at point M and M', respectively, points M and Mf being
infinitesimally close to each other. Denoting

n—1 n —1
= V(k) c an(i MN' = rds + (v^ + dv| c +d cJ ,
k=l k=1 (12.12.23)

we obtain the following formula for the square of the linear element NN

fds + \ ( V(k)d c + c
ds* - dv{
k=1 ^
n—1
ds + ds E V(k) gradll- c + c dv{ (fe)
2 (h - n)
k=l

where the subscript 0 implies that the value is taken on the


supporting trajectory.
it
Taking into account the orthogonality of the unit vectors r
and c, we obtain
k~ 2
n—1 grad II*
1 + c + dv(l) + ■ ■ ■ + dvfn-l)-
{ 2 (h - n) (12.12.24)
k=1

In the particular case n — 3 we introduce the unit vectors n and


b of the first and second normals to the trajectory,
respectively. Since
L - i ~
c = ncos^ + bsin^, c = — nsinip + bcos^
■ 1 “ grad -2-
IT c grad IT c
+
1
^(2) 2 (h - n)
to

2
+ dvf1} + dvf2).

(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 perturbations 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 trajectory.
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 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.

(12.13.1)

The variational equations similar to eq. (11.14.17) are


bap X 4-R18ppT1T(3X6 = 0, (12.13.2)
where R18(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(i)
= &k(i) = 0 ineq. (11.15.27). The result is the following
equations
12.13 Perturbation of trajectories 749

where s and v are components of the perturbation vector along


directions
r and c. The latter is the only unit vector in i7* which is
orthogonal to T. We limit our consideration to perturbations
which are normal to the trajectory, that is we take e — 0. The
perturbation vector then takes the form
l
P = v c, (12.13.4)
where v denotes ’’length” in the metric of the element of the
action as c is the unit vector in this metric. In the metric of the
kinematic element, i.e. the surface R2 (but not R%\) with the
linear element ds, the length of the perturbation vector is
calculated due to the formula
v v
(12.13.5)
y/Hh-Il)
see eq. (12.2).
In eq. (3) the Gaussian curvature of the manifold R% is
denoted by K. By virtue of eq. (11.15.26) it is equal to

K = ^11^22R1212 (12.13.6)

^12
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 cj(i) = 0,o;(2) = 0 in eqs. (11.16.8)-
(11.16.10). Under the conservative perturbation, i.e. Sh = 0, we
d2V( 1) + K ^c, + K ^c, V(2)=
da2 0, (12.13.7)
d V( 2) + K (c,c) z'(i) +K (c,c)
2
=
da2 0,
where and are the components of the perturbations vector
1 2
P = c r>(i)+ c t) (12.13.8)
(2).
750 12. Variational principles in mechanics
12.
Here c and c denote the unit vectors of the normals translated
in parallel
along the supporting trajectory. The coefficients of and z/( 2) in
eq.
(7) are determined due to the rules (11.15.21) and (11.15.23)
and are the
12
quadratic forms of the components of vectors c and c with the
ask = 0 (s ^ k), ass = 1 (s, fc = 1,2,3).
(12.13.9)
[1,1;
In =-nx, of[1,2;
the1]metric 1] = -ny, of[1,3;
the element the 1] = -uz, we[2,2;
action have 1]
due=to eq.
[3,3; 1] = Hx, [2, 2; 2] == [2,3;2] = -n*, IIX [2,1; 2] =
[3,3; 2] = Ily, -ny, [1)1)2] = [3,3; 3] = — —II; [3, 1; 3] =
[3, 2; 3] = — 11^, [i,i; 3] = nz, —II;
11^, [2,2;3]=n*, (12.13.10)
n„
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
2 + BSS, Ask = Bsk,
ss
T = (h-Tt) An + (grad n)
8 (h- n) 4
(12.13.11)

where A is the Laplace operator and Bsk denote the


contravariant components of the auxiliary symmetric tensor of
second rank. They are given
by
(12.13.12)
8 (ft — n)4
(h — n) iixx + —
n) nxy + —nxna (h — n) nX2 + —nxn^ \
2
(h-u)uyy + lul 3,
(h n)uyz + 2uyuz

V symmetric (h- n)n« + |n2


The invariant form of tensor B is as
follows
B = —- grad gradn, (12.13.13)
16 n grad n8(h-ny
(h - ny
3 grad 1
where grad n grad n implies the dyadic product and grad grad
n means the formal dyadic product of the symbolic vector
grad and gradn.
(h — II) All + (gradll)2
(h — II) All + (gradn)2 12.13 Perturbation of trajectories 751
12 12
The covariant + Bcomponents
x
» C Cm, > > cA, c\ of the unit vectors c and c
A

in thesystem
dinate coor xa of the element of the action meet the
conditions
1l ^—> i k3 (i = k) (12.13.14)
2 (z,-TT) ^ CaCa= 0 (i^k) ’

which follow from eqs. (12.4) and (9). By virtue of eqs.


(11.15.21) and
(11.15.23) we find
K (i,i)=
4(ft-n)J
K (-)=
4(/i-n)3

(12.13.15)
The further transformation is based on change from the
Cartesian coordinate 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
^ \ ,, k c • grad II ) +
k Bx» 8(h-ny
1 /c it
c • grad grad
4 (A- nr II- c (12.13.16)
12 3 1 2^
x
-B ^ C\C\= - - -^-3 c • grad II c • grad
11+ s(h-uy 1 2
1 c • grad grad II-
4(h-uy c
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 m
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
i1
arc s and the distances z+i),z>(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
Z'(l) on z>(2) dU
fti 1, /^2 — 1, h% — 1 + 2 (h — II) dz>(i) +
' 2 (h - II)
<9Z>(2)_
752 12. Variational principles in mechanics

hence
_ an i an 2 an
1TT Hll = —
grad 11 = T-——+ c —- - -h c C
h3ds ' ~ di>(i) ' ~ di>(2) ’ ^ di'(fc)
Tensor grad grad II is represented in the form of
2
.dI +
d
c .an i an 2
an products
dyadic
T-— ---V
hsds dv (i) dvd(2) T-——+
hsds dv.—-----h
C (i) Cdv,
—-
(2)
then
i a .an i an 2 an
c • grad grad ndv= (k) r-—-—h c —-----h dv{
cft3as dv, (i) (2)
12
and as the unit vectors r, c, c are independent of we can
conclude that
£ , ,ni o2n
c • grad grad II*dv(k)dvU)'
The Laplace operator is constructed due to the well-
An formulaea ft2ft3 an a ft3fti an a/nft2an
known
hih21hs a*>(1) fti a*>(i) af>(2) h2 dv,(2) a* ft3 a*
and, for v^ — z>(2) = 0, is equal to
d 2n d 2n d 2n
(An)0 = d£,
h + ^(2) + a*2 +
an an
2(ft-n) \dv + 2(h-U) \dv
"(i) "(2) J 0
Inserting these formulae into eqs. (15) and (16) we arrive at the expressions
d2n d2n
K + 2 +
A{h-ny \di>z{1) ds
I 3 / dU
2 V^(i) + ds
4(/i-n) d
J)o
\1
a2n a2n
K(l,l) = 4(/i-n)2 \^2) + as2 +
I 3 / dU 22
2 + a*
4(/i-n)3 \dz>(2) J/0
1 an2 an an
K[ C,C) = +
4 (ft - n) dv{l)dv{2) 8 (ft - n) dv{ i) a*>(2)
2 3
J0
(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 [(/I - H) (nM + n„„) +
m nl + ng] 1 .

(12.13.18)
l4 fa-11) Jo

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 - n) x" - x'2nx + (y12 + z'2) Ux - 2x'y'Uv - 2x'z'Uz = 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
2(h-n) (x'2 + y'2 +h zn,2+^j = 1. (12.14.2)
£^h -n ^+ ntur°-0
i^ - ^' 2(tnr '
(12.14.3)

In the case of a homogeneous gravitational field we can


restrict our analysis to planar motion. Directing axis y along the
upward vertical we arrive at two equations
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
C
x = , ~rwd£Vy2(h- gy) -c2 + = 0, (12.14.5)
2 {h-gy)' dX ~ " “ ' 2 (h-gy)
where e = sign?/' = ±1. Using the notation
V = e\j2 (h-gy) -C
2
, the (12.14.6)
second equation in (5) takes the form
— — (n + C ) dri = da,
2 2
9 (12.14.7)

which can be easily integrated. Thus, Lagrange’s action is set as follows


^ (v
2
A = a = ^{r]0-T]) + VoV + Vo) ~ (12.14.8)

Here 7j0 denotes the initial value of rj


Vo = £o \/2(h- gy0) -C2, e0 = sign y'0. (12.14.9)
Let us take yf0 > 0, then £Q = e = 1 on the upward part of the
trajectory, on which y increases from yo to y*
2 h-C
2
y* 29
corresponding to r]0 = 0. On the downward part y decreases
from yo to y* and £ — — 1.
The constant C can be expressed in terms of the initial value
XQ of the derivative x'. Replacing then da in eq. (5) by its value
(7) we come, after integration, to the trajectory equation

X = x0 + 2 (ho - gyo) y (Vo ~ v) ■ (12.14.10)

Hence, on the upward part


X = Xo + 2 (h0 - gyo) y ^2(h-gy0)-4:(h- gy0)2 x\

\J2 (h - gy) - 4 (h - gy0f (12.14.11)

at the turning point of the trajectory

X* =x0 + 2(h- gyo) —\j2{h- gy0) -4 (h - gyo)2 a# (12.14.12)


12.14 Examples 755
and on the downward part
x' \]2{h- gy0) - 4 (ft - gy0 f xfi
X = x0 + 2 (ho - gyo)
^2{h-gy)-A{h- gy0f x'Q2
(12.14.13)

For a given h the solution has three constants, namely


and describes a bundle of trajectories with the turning
xo,yo,xf0
point at point xo,?/o- Any particular curve C of the bundle is
specified by slope x'0. On a curve Cf of this bundle, which is
infinitesimally close to curve C, the slope is XQ + 6x'0 and
variation determining the transition from C' to C, is given by
<Sx,

Sx - & -6x
c
dx '
~ dx'0dx°- (12.14.14)
Prom eqs. (11)-(13) we obtain
dx = x - x0 4 (ft - gy0) Xq
dx'0 x'0 7]Q \f)\ (12.14.15)

where the positive and negative signs correspond to the


upward and downward parts, respectively. At the turning point
of the trajectory
dx X -jf. 2 (h - gyp) xfi
dx' (12.14.16)
0
XQ 1-2 (ft - gy0) x'q
Xn
It follows from eqs. (14) and (15) that variation 6x reduces to
zero for non-zero 6x'0 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 = x\ and y = yi
= \ 2 (ft - gy0) Xp _ ft /ft Vo
2/1
) 1 - 2 (ft - gyo) x'Q g
V0 2
9 \9 \g
T _r. i [2 (h — 5fto)]
3/2
X'Q _ 2(h-gy0)x'0
*
9
\/l -2 (ft - 5J/o) xo 9 Vo
,
(12.14.
17)
Thus, the infinitesimally close isoenergetic trajectories
intersect at point
(xi, 2/1) which is the kinetic focus corresponding to the initial
point (x0, 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.
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
x\ = x0 + — cot a, 9 (12.14.18)

Excluding a from these we obtain the equation for the locus of


the kinetic foci corresponding to the initial point
vl g (X! - x0) h_ _ gQei - x0)2
yi - yo = -x- -
2g
(12.14.19)
g 4()‘
Vo

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 (x0, 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
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

nx = nr- TT _ Hr n X2 X2
llxx — T IIt’7’ n 117*
Q•
12.14 Examples 757

By virtue of eq. (13.18) we find


1 h-U d
+ n* (12.14.20)
K = 4(h-U)3 r dr
rllr

For instance, we can take


II (r) = firn signn, (12.14.21)
where the factor sign n is entered to ensure that the force is
attractive for fi> 0. Then we have
2 2
hii\n\nrn 2hji\n\nrn
(12.14.22)
~~ 4(/i-n)3 ~ v
5

The necessary condition for stability of the orbit is the


positiveness of hn. In particular, for the circular orbit

— = nr n—1 -firn (n + 2) signn,


r * = T + n:
so that
hn — - \n\ iirn (n + 2)
the above criterion leads to the inequality n > —2. The case n =
— 1 corresponds to Newton’s attraction force, then
5 h
= -h
v2 r 2r
and due to eq. (13.3)
= v o cos —— + Vn JTn sin ——.
v (12.14.23)
y/rji Vn1
It is easy to obtain the dependences versus time

2, + v'0 \l— sin


G — V t — -, V = Vo COS (12.14.24)

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 2p and
A A 71-
/i < A, A + p < — •
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 ( O B ) generating
lines. The equations for the perturbed trajectories are
required.
The position of point M on the conical surface is described
by the distance l = O M from the vertex measured along the
generating line and angle ( p between the vertical plane
containing and the plane passing through the point and the
axis of the cone, see Fig. 12.8.
\
n = — gz = —gl (cos fi cos A + sin fi sin A cos <p). J
In this case h = 0 and the square of the element of the action
da2 = —2Tlds2 = is
-211 ( dl +1 sin /juhp ). (12.14.26)
2 2 2

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 /,</?

g
= — — cot /I (cos A sin p — sin A
cos fi cos p)
12.14 Examples 759

The value of K should be determined on the supporting


trajectories, i.e. at p — 0 and p = TT. For these trajectories
—II = gl cos (A =F AO >
All = =by cot g sin (A =F AO > (grad II)2 = g2 cos2
(A =p g) and substituting into eq. (13.18) yields
K = ±_____________________,
4gl3 sin g cos2 (A =F /i) ’
where the upper (lower) sign corresponds to the lower (upper)
generating line of the cone. Differential equation (3) takes the
form
d 2v sin A v—0
2
da sin/i cos2 (A =F g)
4gl3

and it remains to express l in the supporting motion in terms of


the independent variable a. By virtue of (26) we obtain
i
9a2
J V^mdl = |^/2gcos (AT/I)I3/2, l 3 8g cos (A
0 =p AO
and furthermore
d2v 2 sin A v
n (12.14.27)
da2 9 cos (A =p g) si g v
2

We arrive at the equation of Euler’s type which is integrable in


quadrature. The solution has the form

i/ = (CKT* + C2a~
q
), 2 sin A
1
Q =
4 ^ 9 sin g cos (A
=p g)
The value of v increases beyond all bounds as a 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 a —► oc and the necessary criterion of
stability is also not satisfied.
The value of v, which is the ’’length” in the metric of the
element of action, is related to the normal deviation
v — l sin g6(p
from the supporting trajectory by eq. (13.5). Hence
v 2v 2 1 (Ckt9 + C <r-q) •
2
l sin fiy/2gl cos(A=FaO 3crsin//
3sinn y/a
On the lower generating line q < — and Sip —> 0 as a —> oc.
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 F\. The distance
FFi = 2b and line FF\ 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 ^ due to
eq. (12.28)
i~
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 and v^) from the supporting trajectory are
n
II = 2fi (a2 + b2) 2 signn

and does not depend upon the position of the particle on the
trajectory as it is independent of a. For small deviations

II = /x (r" + r2) signn = /z | (a-i>(1))2+ ( b - V ( 2 ) ) +


R >j
(a-j>(i))2+(6 + i>(2))2] 2 > signn
‘ n

= n (a
2
+ b2 — 2az>(!) — 26P(2) + ^\i) T ^(2)^ T

(a2 + b2 — 2az>(i) + 2bu^2) + ^\i) T ^(2)^ sign n,

It is sufficient to retain the quadratic terms in the expansion of


II in series in terms of P^and z>(2). Then we have

n= 2/i (a2 + b2) 2 nau{1) n *>(i) + f>\2)


a +b 2 2
2 a2 + b
2

a *>m +& ^(2)


2 2

n (n — 2) (aTi)2 + b2)2 signn (12.14.28)


K (c, c) a (n +42) +
2
4ft 2 n 2
2

2fi \n\ a (a + 6 ) ^
2

K (c,c) a2 + b2 (n — 1)
2(i \n\ a (a + 6 ) ^
4 2 2 n 2

The motion along the supporting trajectory is stable under the


following conditions
a2 a2
-
n> 2 + 4
52’ (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.
R — bn (12.15.1)
762 12 . Variational principles in mechanics

The generalised coordinates q2,... , q n are assumed to be


chosen such that their zero values correspond to the
supporting motion. These coordinates and their derivatives
q2,... , q'n are small along the perturbed trajectories. Hence,
only the terms linear in #2, • • • , q n, q^ • • • ? q'n should be kept
in the differential equations of the perturbed trajectories which
are the variational equations. This allows us to replace
coefficients b k by their expansion in series in terms of q2, . . .
s

, q n , q 2 , . . . ,q'n- The series for bn should be expanded up to the


second order terms, the series for b\ up to the first order terms
s

and it is sufficient to take the zero-order terms in the


expansion for bsk when s > 2 and k > 2.
Within this accuracy we construct the expression for y/R
Vr= Vhi + Yl blsblk
s=2 bni/bn QsQk *
(12.15.2
)
The zero indicates that the value in parentheses is calculated
for zero values of q2,... ,q . n

Let us proceed to calculating the derivatives of v/7i*


appearing in the Eulerian operator. It is necessary to bear in
mind that coefficients bsk depend upon q\ both explicitly and in
terms of q , ■. ■ ,q , the latter being the sought-for functions of q\.
2 n

Then we have for .s = 2, ./ . .b ,k_______bigbik


di/R n
\
dq's bis
y/bii V V ^ i i
s
bii\/bu)i Ik>
d dR
d bu d bis \
dqi dq's dqi V h i + ^ 2 q k {dqk Vb^)0 +

f bsk bigbik \ „ ^ f blsblk \


k^k V v ^ n bny/bn)o
E2 bsk
k=

dVR djbTi _ \VbTi bnVbTi)~


k= 2
+
d bik_\
qk

dqs dqs dqs y/bn J o


We arrive at the variational system of linear differential equations

bisbik d bu _ d bik \
biii/bii dqkVhi dqs VbiiJ q'k
0
, 9 bis _ 9Vbn . 9Qi
d / bsk
bisbik \ f dqi \\/hn =
Vhi dqs (12.15.3)
bn^/bii)oqk 0.
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 0( in the supporting motion.
In the perturbed motion the body position is determined by
Euler’s angles ip, $, ip, angle $ being small. The projections of
the angular velocity vector of the body onto the axis bound to
the body are
UJ = 'd cos <p + ip sin $ sin ip, ujy = —P) sin <p + ip sin $ cos ip,uoz =
X
ip -f ip cos d.

Zhukovsky [97] suggested new variables which allow the


T — - (ACJI + Buj y + CLU z)
2 2

in the case of small d. Assuming


& = tp + ip,
we can recast the formulae for projections of the angular
velocity in the form

Introducing the new variables which are projections of the unit


vector of axis 0( on axes Ox and Oy
x = i? sin y — d cos ip,
we obtain
xy — yx ipd2 = xy — yx.
tany? = 5—
y cos^ ip
Hence
wx — y + wy = —x + y<&, Luz = <&

and the expression for the kinetic energy takes the form
T= [C + (A- C) x2 + (B- C) y2 + (2A - C) xy'+
(C - 2B) x'y + Ay'2 + Bx'2} ,
764 12. Variational principles in mechanics

where a prime denotes a derivative with respect to <f>. The


potential energy is given by
n = QlcosD = Ql[l- y) = Q l ( l -

where Q and / 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
h = ^C$2 + Ql,

where 4>o = { w z ) o - This value of the energy constant is also


held in the case of conservative perturbation. Then we have
2 (h - n) = C$20 + Ql (x2 + y2),
and due to eq. (1) we obtain the following expression for
function R
R= C$1 + Ql (x2 + y2)] [C + (A - C) x2 + (B - C) y2+
(2A - C) xy' + {C- 2B) yx' + Ay'2 + Bx'2 j .

Therefore, if we ascribe to the coordinates indices 2 and 3,


respectively, we obtain, with the required accuracy, that
Ql 4-C | o 2
611 = c2$2 1 + + ^7— x + , (-2L + B-C
C$20 c) \c$2 C
bi2
vu- = ±$o(C-2B)y, -^= = 1$0(2 A-C)x,

^22 ^33
— ^0^5 ^23 = 0, = $0 c.
(12.15.5)

Canceling out the multiplier <j>0 we can set the equations of


motion (3) as follows

Bx" + (C-B-A)y'-(^+A-CSjx = 0,
(12.15.6)
Ay" + (C - B - A)x' - B - C^J y =

0.
Looking for the particular solution in the form
x = M sin (A4> + a ) , y = N cos (A4> + a ) ,
12.16 Hamilton’s characteristic function 765

we come to the characteristic equation for A


2
B X 2 + S I + A - C ] (A\+ % + B-C) -{C-B-Af A2
$1 / vn
AB + (A- C) { B - C ) - { A + B)%
= ABX
4
- $oJ A2 +
Ql
+ A - c ] ( S i + B - C ) = 0.
n
Hence, the
necessary condition for stability of the vertical
position of the axis is positiveness of the roots A^A^ of this
equation. This yields the following criterion
b = AB + { A - C ) { B - C ) - { A + B ) % > 0,

Ql_+ B - C >0 (12.15.7)


2
b - 4ABc > 0.
Let 1 = 0 which corresponds to the case of rotation about the
immovable centre of inertia. The characteristic equation has
the roots
A? = 1, A 1 = Jb(B-C)(A-C),
and rotation about the axis of the maximum or minimum
moment of inertia is stable.
For A = B and l > 0 we arrive at the condition for stability of
the vertical spinning top

4 AQl
*0>
~Cr’
This condition also follows from formula (7.9.28) is we take A2
= C'2 — 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 constraints 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
dS dS dS (12.16.1)
_ + Hlq 1 ,<7n = 0,
dt ’ dqi ’' *' ’ dqn
766 12 . Variational principles in mechanics

depending on time, the initial and actual values of the


generalised coordinates as,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 equations do not contain t
explicitly and do not change under the replacement of t by t —
to. Hence,
S S , • • • 5 qn•> OLi , . . . , (y.n,t £0) •
(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 ( q i , . . . ,qn,'y1,... , 7 n _ i , f t ) + 7„ where 7n is an

additive constant. After a proper choice of ryn we obtain


V = - h ( t - t 0 ) + W ( q u . . . ,gn,7i , . . . ,7„-i,h) -
W ( < * 1, . . . ,an,'y1,... ,ln-i,h) •
(12.16.3)
In order to determine the constants j1 , . . . , 7n _ x and time t — to
we can construct the following n equations with the help of eq.
(8.14)
ftj^
1 'In— 1 5 ^o) \yv (<24 5 • • • 5 Q.m 7l 1 • • •
W ( c * i , . . . . a n j !, . . . =0,
dv 9 rw(
0^ = 0^^ 0*1’-" ,In-1,h)~ (12.16.4)
h
W (ai,... ,ln-i, )} = 0
(k = 1, . . . ,n - 1) . ,

From these equations we find the constants 7l 5 . . . , 7n _ l 5 ^ which,


by substituting into eq. (3), yield the principal Hamiltonian
function. But we can also proceed differently than above.
From eq. (4) we can express t — to, 7i , . . . , 7n — l m terms of a q , . . .
, a n , h and then exclude them from eq.
(3) . The result is
S + h(t-t0) = W(qu... ,gn,7i,... ,7n_i,h) -
W (oil, . * * , Qn, 7l> • • • , In — 1’ (qi, • • • , qn, Ql, • • • ,
Qn, ,
(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

{^-^Wl — Wn)=K
H (12 16 6)
- '
in which the constants a q , . . . , a n denote the initial values of the
generalised coordinates.
12.16 Hamilton’s characteristic function 767

The function
A ((/l , . . . , Qni OL1 5 • • • 5 ^715 S (Ql 5 • • * 5 Qfl
•> OL \, . . . , , t ^q) T /l (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 Lagrange’s action. Indeed, by virtue of
eqs. (8.1) and (8.2) and when the system possesses the
energy integral
we have
S = J Ldt = J (T-U)dt = J 2 Tdt - h(t - t0) = A - h(t - t0) .
to to to
(12.16.
8)

Casting here Lagrange’s action in the form of eq. (11.3) or (11.4)


(i) 1/2 Q l
2(ft-n)EE Ctsk^Qs d*Qk = J y/Rdqi (12.16.9)
A= /
(0) s=l k=1
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 construct Hamilton’s characteristic function. On inserting the
general solution of Jacobi’s differential equations (11.6)
qs = q s ( q i , c 1 , . . . , c2n_2) ( s = 2,... ,n). (12.16.10)
into the following expression
Qi

Oil
we can express the constants C \ , . . . , c2n_2 in terms of q \ , . . . , q n ,
oli , • • • , an. To this end, we should use both equations (10)
and the initial conditions
a s = q s ( a 1 , c 1 , . . . ,c2n-2) (s = 2 , . . . , n ) . (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
768 12. Variational principles in mechanics

relationship (5). The general case of asynchronous variation is


implied in which the two positions of the system
{ & ( * ) } a n d { q * s ( t + At) = qs ( t ) + Aqs}

are compared at different values ft and ft + 6h of the energy


constant, notation Aqs being given by formula (9.4).
We obtain
A A = V — A qdqs dA \ DA
7^1 \
+
•7-—e Sas ] + -77-
ftft
(dS
oadS s J oh
\ dS A . \ ci 7 A
d ^ s ) ~di - 6 h + h A t >
+ 6as + At+ to)
=E

where by virtue of eq. (9.4)


^dA. ^dAs A y^OA.

dA • 9 A
53 -fa *
6 3 =
53 ~0n~6^s ^A t
8=1 OQs 8=1 OQs
and by
A.
At ,TA ft At T y —dS
analogy —AAqs
dt t=i ^
d
dS
^dSc 6(dS Aas. \ ^OS c ,r ,,A
— 2_^ o *s + ( "o7 + 53 —*8 + ^ ) At — >3 o—+ (L + ft) Aft
v^^^ / ft? ^
Relationship (12) takes the form

'dA
(2 T-L- h) At + dh (t - t0) 6h = 0.

The coefficient of At disappears by virtue of the energy integral


(7). The first equation in (4) and eq. (5) yield
dA
dh = t-t0. (12.16.13)
It remains to equate the coefficients of the independent
variables 6 q s and S a s to zero. Then recalling eq. (8.8) we arrive
at the
dA
following
dS
relationships
dA d S = -(3 , (s = 1,... , n),
dqs dqs = Ps, d a s das s (12.16.14)
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 function answers the question as to what initial momenta were applied to the
material system under a prescribed energy constant /i, 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 q\ , . . . ,qn can be expressed in terms of the
constants a q , . . . , an,i , . . . , /3n, h. However it is not possible since
these constants enable the n - 1 coordinates , <7n to be
expressed in terms of q\.
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 metric of the kinematic element
2Tdt2, we can write the first set of equations in eq. (14) in the
v = grad A.
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 determined 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 surfaces 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
(2)
J n ( x , y , z ) ds,
(i)
where n denotes the refractive index of the inhomogeneous
isotropic medium under consideration. Comparison with eq.
(10.7) shows that the corresponding dynamical quantity is the
value of the momentum
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
770 12. Variational principles in mechanics

energy
n2
n = h- 2 m

Details of the optical analogy that played an important part in


the development of wave mechanics can be found for example
in [95] and [49].
Let us turn to examples of constructing the characteristic
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)

^ ( V 2 + V o V + V o ) + c2
a = - (% - v) (12.16.15)

where, due to eqs. (14.6), (14.9) and (14.10),


(j ___________
x-x0 = — (v0-v), Vo = \/2(h- gy0) - C2. (12.16.16)
V = \/2{h - g y o ) - C 2 .
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 ( x o , y o ) and the final (x\,yi)
positions of the particle and the energy constant h.
The calculation is as follows, cf. [85]. We have
2(y- yo) C
Vo ~ V = 2 . g { y - y o ) , Vo + V X — Xq
and we use the identity

2 , 2'no (Vo + vf + ( V o - V?\


+ T

to come to the following equation for C 2

2h-g(y + yo)-C2= C2 + ^ (^~2'Xo) • (12.16.17)


X — Xq
Denoting for brevity

M = 2h- g(y + y0), N2 = g2 \(x - x0)2 + (y - yo)2] , (2.16.18)


12.16 Hamilton’s characteristic function
771
we obtain from eq. (17)
C2 M± y/M2 - N2 1
(y/M + N ±y/M - N) ,
g (x - x0)
2 2N2 AN2
g2 (x - = 2 (M =F VM2 - W2) = (VM +A^zp y'M-Af)5
xq)2
C2 (12.16.19)
Casting now eq. (15) in the form

A = <7 = - (ri0 - ■^(vo + v) + — (g0-v) +C2


rj)
X=— Xo y-yo C2 +
1 g2(x- x0)
+ C2
X— 12 C2
C Xo
= i [y/M A- N T y/M -N^ \(M± y/M2 - N2) +

1 (M =F x/M2 - N2}

= T (x/M + N =F y/M — iv) (2Af ± x/M2 - ./V2) ,

we come to the following expression for the characteristic

function _1 3/2
=F(M-AO3/2 .
(M + N)
_■
3g .
At the initial point of the trajectory N = 0, taking the plus sign
in the latter equation, we have A = 0. 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
3 g (M + Nf - (M - A0
/2 3/2
A
(12.16.20)

and after this


A
3(7 (M + N)
3/2
+ (M - Nf/2 (12.16.21)

In accordance with eq. (13) we have


d - (x/M + AT T y/M- N^j = t
A — h. (12.16.22)
dh
This is the equation of the curve which contains all the
particles emanating
772 12. Variational principles in mechanics

By virtue of eq. (14) we find the initial momenta


^ [VmTN±VM-N\ (x - xo) = -xo,
QA 1 _____ ________
— = WmTnt>/m-n\ + (12.16.23)
[JmTN±VM=N] (y - yo) = -yo,

Each of these equations represents the equation of a


bundle of the isoenergetic trajectories originating at point
(xo,2/o)- Each curve is specified by the value xo or yo of the
slope. Noticing that

we arrive at the following notation of equalities (23)


1
x — • u * \ , y ~ yo •
Xo t - so, -g {t - to) + -f- to = V°'
t0
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 (r0, ip0) of the trajectory has
the form
W = f3v(<p-v0) + J ‘\J2hr2 + 2fir — (12.16.24)
ro
and equality (4) is as follows
r
dW ■I dr
d(3 = V ~ <A) - A = 0. (12.16.25)
•y^2 hr2+ 2/ir — (3\
v

This is the equation for the bundle of isoenergetic trajectories


originating from point (?'o, <£>0). Using this in expression (24)
leads to the form r
2 hr + 2fi
W ■/ -.dr. (12.16.26)
yj2hr2 + 2 fir — /?;
ro
The characteristic function should be obtained by removing
(3^ from equalities (25) and (26). In order to simplify the
further equations let us introduce the notation
2h =
~a’ @1 = Va ~ e2) = VP, (12.16.27)
12.16 Hamilton’s characteristic function 773

where a, e and p denote the major semi-axis, the eccentricity


and the parameter of the elliptic orbit, respectively.
Then we can write

2hr2 + 2fir — /?2 = — [2ar — r2 — a2 (l — e2)] .


Introducing the notation

y/2ar — r2 — a2 (1 — e2) = £, yj2aro — 7g — a2 (1 — e2) = Co

(12.16.28) we arrive at the equalities


e2rr0 cos (p - p0) = (l - e2) CCo + (a - r — ae2) (a — r0 — ae2) ,
(12.16.29)

w= - arccos -T [CC0 + (a - r) (a - r0)] 1.


(12.16.30)
I CL ae I

The identities

a (l — e2) — r = \Jr2e2 — (1 — e2)


(12.16.31)
£ 2,

follow afrom
(i - eformulae
2
) - r0 = v/rge
(28). “By
2
(1using
- these, eq. (29) has the
form
e2rr0 cos (<p - <p0) - y/r2e2 - (1 - e2)C2\/rge2 - (1 - e2)Co = (l - e2)
CCo-
Taking the square of both sides, cancelling out the factor e 2
and applying identities (31) we come to the relationship
e2r2rl [l + cos2 (p — p0)] — 2rro cos (p — p0) [a (l — e2) — r] x
[a (1 - e2) - r0] - (l - e2) (r2(20 + r02C2) = 0,

(12.16.32)
which is a quadratic equation for p. It determines the value of p
1 1 2
F(p) / P2 +
r, X
12 + ^2 ~ — COS
VP ~
^0-) pQ)
1 + cos (p 2[ i + - [l-cos(p-p0)]p +
. r r0
[1 — cos (p — p0)]2 = 0. (12.16.33)
774 12. Variational principles in mechanics

We find two solutions


1 - cos (ip - ip0) 1 + cos (<p - p0) 1 + J_±
Pi,2
2 cos ((p-ip0) _ 2a r ro
r2 r2
rr0

(tp — 'l + cos(y> - <p0) 12 (12.16.34)


+ COS
______________
¥?o)l
ar0 rr0
1_
corresponding to two curves joining these points. At the
culminating point (ri, Lpx) 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 —
ro) (12.16.35)
ri r o 4aCOS
- r0 (<£>! - ip0)
4a - r0
As ro < 2a and
0 < e = 4aro- r0 < 1,

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
m 4a (2a —
2ae = 2e-+ Pi ro) 4a -
1 — £z
where a and pi denote the major semi-axis r0 ’ and the
parameter, respectively. Inserting the expression for e we find
that the focal distance equals 2ae = ro, the second focus of
ellipse (35) is located at the initial point and the major axis 2a
is equal to
2a = -j =4 a-r0.
(12.16.36)
The elliptic trajectory and ellipse (35) are marked in Fig.
12.9 by C and E, respectively.
Let F\ 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
M0Fi + M0F2 = To + M0F2 = 2a, M*Fi + M^i^ = T\ + M*F2 =
M0F2 + M*F2 = 4a — ro — r\. (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 M0 we have

M*F\ + M*Mo = ri + M*Mo = 2a, M* Mo = 2a — ri = 4a — ro —


ri, and a comparison with eq. (37) yields
MqF2 + F2M* = M*Mq.

If follows from this result that line segment M0M* 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.
T)
(34) into eq. (30) and replacing e2 by 1 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
_ dA dA
(12.16.38)
2a2
and to determine the actual
0
dhand initial
da p,values of momenta we
apply formulae (14)
P<p = dA__9A_ _dA _ dA
(12.16.39)
dip dp0 ’ dr ’ r dr0 ’
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 i?* 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)0 and consider a family of
hypersurfaces A(qsHamilton’s characteristic function being
constant on each of these surfaces. All trajectories belonging
to the set oon_1 trajectories of the bundle with vertex at the
initial position (gs)0 intersect orthogonally each hypersurface
at each point. Let us consider the points of intersection of the
trajectories with a certain hypersurface A = Ao and lay off the
segments of the length do from the intersection points. The
loci of the ends of the segments are the hypersurfaces of the
same family since the segment of the arc of geodesic i?* 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 — l)-
dimensional manifold is described by (n — 1) parameter x1,... ,
xn~l. We can introduce the system of coordinates x1,... , xn_1,
n—1 n—In—1
ds2 = gskdxsdxk = gnndxndxn + 2 gindxldxn + EE gskdxsdxk.
i=1 s=1 k—1
(12.17.
1)
On the hypersurface A = const, i.e. where dx1 = ... = dxn~x =0
we have
ds2 = da2 = gnndxndxn
and thus
However
9in — 0 — 1,...,72 1) (12.17.3)
12.17 On the character of the extremum of Lagrange’s
777 action

which follows from the orthogonality of the trajectory tangent r


to any infinitesimal vector <5r on the surface A = const
passing through the point under consideration and the
definition gin as the scalar product of the basis vectors and r.
Hence, on an arbitrary curve of manifold i7* we have
n—1n—1
ds2 = da2 + ^2 9skdxsdxk,
where
n—1n—1
EL gskdxsdxk
s=1 fc=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
(i) (i) n—1n—1 1/2 (1)
!H > J dcr =
(i) “ (o) (0) (0)
a
5 5
da + EE gskdx dx
2 s k (0)
s—l k—1 <r(1) (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)
reducing to zero at a = do- The solution of differential equation
(5) is called
non-oscillating in the interval (cri,^) if it reduces to zero not
778 12. Variational principles in mechanics

once. By virtue of the theorem which states that if


at <Ji < <j < <J2 K < 0,
(12.17.6)
then the solution of differential equation (5) is non-oscillating
in any subinterval (crj, <t2) of the interval (cr 1, <72). This
means that there exists no perturbed trajectory Cf which is
infinitesimally close to Co, begins at point 0 = <Jo (<Ji < <r 0 <
a2) and intersects C within this interval. Thus, this interval
contains no kinetic foci and Lagrange’s action along the
trajectory (<Jii<J2) 1S minimum.
For example, let K = —m2 = const, then the boundaries a 1
and 02 move to —00 and cx), and the interval (crj, 02) without
kinetic foci can be chosen such that within this interval the
absolute value of the solution
z/
v — — sinh (a — <To)
(12.17.7)
m
can be made as small as is wished. This can always be
realised by a proper choice of vr0. This ensures that C'
remains a trajectory which is adjacent to C and does not
intersect C since function (7) does not vanish at o ^ ao-
Let us assume that
at <ti < a < <72 i^max > K > Km[n > 0.
(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 7r/y/Km[n and greater than 7T/y/Xmax • If we take
the solution which turns to zero at a = Jo we can assert that
interval (<Ji,<T2) of trajectory Co has the conjugate kinetic
focus provided that
02 > —== o\ < o < <t2.
(12.17.9)
hence,
v
h — gy = g (y — yo) + g(y -y0) <h- gy0,

(12.17.10)
where the equality sign corresponds to the vertical motion at
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
(#i - xq)2-\-{y\ - yo)2 = 4 h
9 2/o (2/i - 2/o)+(2/i - VoY
12
(2/1 - (12.17.11)
2/o)
Taking into account inequality (10) it is easy to conclude that
the required minimum is reached at y = y\ corresponding to
the equality sign in inequality (10). It follows from eq. (14.19)
that x\ — x$. 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
X = X0, y = -=yi
(12.17.12)
9
into formula (16.20), we find the constant in the equation for
the curve A = const
{[2ft — 9 (2/1 + 2/0) + </12/1 -2/o|]3/2

- [2 h-g (yi + 2/o) - 9 \yi ~ 2/o|]3/2| = [2 ~ 99o)]3/2 ■

The obtained curve


3/2

A= 2h - g (y + y0) +g\l{x- ^o)2 + (y ~ Vof


39
3/2 '

g(yi+ 2/0) - g\f{x - x0f + (y- yoY


2h ■

^ [2 (h- gyo )}3/2 (12.17.13)

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
A = 7 at 7 < 7* = L [2 (h - gyo)f/2 (12.17.14)

are also closed and lie in domain G. They contract to the


bundle origin as 7 decreases, and the ’’curve” 7 = 0
degenerates to point xo,2/o- 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 = 7 for 7 > 7*. Let us fix
point (xj, y\) of the safety parabola, then making use of eqs.
(14.19) we find that
7 = 4- [4/i - 2g (yl + 2/o)]3/2 > *•

The required equations have the form

= ^ [4h - 2 9(2/1 + yo)) 3 / 2 ■ (12.17.15)

The upper and lower signs correspond to the branches A\


and A2 shown in Fig. 12.10. This figure displays two
trajectories (7* tangent to the safety parabola at points
(=bx\,y\). Let us agree to categorise the bundle trajectories as
being ”steep” and ’’flat”. Any point of the domain bounded by
curves (7* and A\ can be connected by a flat trajectory, for
which Lagrange’s action is minimum. However any point of
the domain bounded by the safety parabola and curve A\ is
reached by a high trajectory orthogonal to branch A 2. The
action along the path of this trajectory containing the kinetic
an a 12 ••• din
&21 <222 ••• d2n

^ml dm2 ••• dmn

which contains m rows and n columns is referred to as a m x


n rectangular matrix. The first and the second subscripts of
the matrix element 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 rather
than with single elements considered separately we use the
an «12 ... a\n
0* «22 ... a2n
21
Q" ••• ^nn
nl
2

(A.1.1)

In order to save space, we will also write


& — ||&ik || 5 ii 1,m\ k i). (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 (z = l,...,m; fc = l,...,ra).
(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
Xi
x2
(A.1.4)

or as a row which can be understood as a 1 x n matrix

x' = \\xi,x2,...,Xn\\. (A.1.5)

2. Having a system of m functions of n variables aq, ...,x n


fr=fr{x 1,.(r = 1,m), (A.
1.6)
we can defineElthe m Of
x n matrix
Dx r
dx , (r= s =
(A.1.7)
which is referred to here as Jacobi’s matrix of this system of
(r = 1,...,n)
••• aai\kt
••• i2kt

a
iski aisk2 ••• aiskt
an
ai2 ••• ^1 n
&21 &22 ^2 n

ttnl an2 ••• C^nn

(A.1.11)

where ii, «2, • • • ,is are any of the numbers 1,2,... , m and Aq,
&2,... > kt are any of the numbers 1,2,... , n. Clearly, s < m, t <
n. Then, the s x t matrix (11) is referred to as the submatrix of
a and is denoted as
list (i
’ i\ii‘ii k Aq, k25 • • • 5 Aq) • (A.l.12)
If we deal with a quadratic submatrix, then s = t. The
determinant, constructed from the elements of submatrix (11),
is referred to as the minor determinant of order s of matrix a.
Clearly, the highest order of the minor determinant ofamxn
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 \a\ or det (a)

\a\ = det (a) (A.l.13)

The determinant of the quadratic n x n matrix is one of the


minor determinants of n — th order, namely that the
determinant for which the
dfi dfi dfi
dxkl dxk2 '
dh dh dh
dxkl dxk2 ‘ " dxkm (A.1.14)

dfm dfm dfm


dxkl dxk2 ' " dxkm
&z+l,/c+l •
ftz+2,/c+l •••
&z+2,/c+t

1 ••• (li+s,k+t
lla.ll(1,1) 1 m^II
II il (s,n—t)
II H(m—
s,t) \ Hall (m-s,n-t)

which enables values X/Cl,..., Xkm to be expressed in terms of


the remaining n — m values Xkm+1,x/^from the following system
of equations
fr (xi,x2, -,xn) = 0 (r = 1, . (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

IN (i+i,fe
+i) (A.1.16)
(s,t)

Then we can represent matrix a in the following way

(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
[ai,ai] ....
[a2,ai] .... [oq,
[a2,aan]n]

[a„, ai] . ...


On\0L
] n,

[/3i,ai] ...• [Pi ,OL ]


[/32,ai]... . [/32,Q;n]n

[/3n.., .OL
[Pm
ai] n\
[ai ,fii] Pn][oil,
...
[a2, fii] ... Pn\[c^2?
[an> fii) ... [o;n,/^n

[fiufii) •Pn\
[/^l •>
■■
E/3
•• 2,/3i] •Pn\
[/^2 ’

[firvfil] -

(A.1.22)
It can be represented by four nx n matrices denoted by [[aa]],
[[a/?]], [[/3a]] and [[fifi]]

A= M \WfiW (A.1.23)
[[/3a]]
[[fifi}}
If we assume that the Jacobian is not equal to zero
D{qi,-,qn,Pl,--;Pn) ^ 0
(A.1.24)
D (ai, - -., an, fiii fin)
786 Appendix A. Elements of the theory of matrices

then system (21) is resolvable for ar,/3r


°^r — OLr ((/I, qn\Pi, ..., pn), / _ \
/3r = 0r (qi,-,qn;Pl,-,Pn) , (
~ (A.1.25)
The following expression

da d(3 da d/3
(/?,<*) (A.1.26)
dqr dqr
dpr dpr
is called Poisson’s brackets of the system of functions a, f3
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 (a*,
a^), (ai,f3k), (/3i? a^), (/3i? /3fe). The notation analogous to eq.
(23) has the form
((aa))
P= (A.1.28)
((«/?))

A.2 Operations on matrices


1. Transposing. Given a m x n matrix, we can determine a
nxm 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
(af)f = a.

Elements an of the square matrix is called the diagonal A


square matrix a is said to be symmetric if aik = n/ci- 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 a^ = — a^.
The examples are Lagrange’s matrix (1.22) and Poisson’s
\a\ = \a'\. (A.2.2)
0 —as a2
a= as 0 ~ai
—a2 a\ 0

(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
lxn 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 6,
respectively, i.e.
c = a±b if cik=aik±bik = k = l,...,n). (A.2.4)

It is evident that
a + 6 = 6 + 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 A is the matrix with the elements equal to Aa^. The
notation is as follows
Aa=||Aaifc|| = k = l,...,n).
(A.2.5)
In particular, for A = — 1 we have the matrix —a with the
elements — a^. Hence, the transpose of the skew-symmetric
matrix yields
For instance,
a! = -a.

(A.2.7)
In accordance with the above definitions, the identities
11.
a{k = 7^ {a%k T aki) T ~ (aik akf) ? (2, k 1,..., nj,
a = - (a + a ) + - (a — a')
f
(A.2.8)
788 Appendix A. Elements of the theory of matrices

and determine the decomposition of the square nxn matrix a


into a symmetric part
2 (a + a!) — - (a' + a) (A.2.9)

and a skew-symmetric part


\ (a ~ a')' = -i (a - a') = -a).
(A.2.10)
Example 3. By virtue of eqs. (1.20) and (1.22), the
transpose of submatrix [[a/?]] of Lagrange’s matrix (1.23) is
given by
[[a/3]}' = ~ Woe}}. (A.2.11)

The analogous formula takes place for the submatrix of


Poisson’s matrix (1.28)

((a/J))'= -((/?«)).

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
v
cik = (i = l, k=l,...,n).
(A.2.12)
r= 1
Clearly, the multiplication of matrices is not, in general,
commutative. For example, if m ^ n the product makes no
sense at all. If m = n, then ab determines a square nxn matrix
whereas ba determines a pxp matrix. Only square matrices
can be commutative. For example, the unit matrix E n is
commutative with any square nxn matrix
aEn = Ena = a.

(ab) c = a (be) = abc. (A.2.14)

The property of matrix multiplication is distributive over


addition, i.e. a(b + c) = ab + ac.
xiyi xiy2 . •• X! yn
X2y X2P2 • • • X2yn
xy =
xnyi xny2 • • • xnyn
a\b\ aib2 aib3 1
&2b a2b2 0263 > . (A.2.21)
\«3^1 03^3 J

(A.2.19)

(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

This table coincides with matrix a'b, thus we can write down
the following equality
ab = ab’ (A.2.22)
0 —a3 a2 bi -0362 +
ab = b2 aa2b3
3bi - axb3
<5
O
C

1
s
=
—a2 a\ 0 bs -a2b\ + a\b2

or
ab = c. (A.2.25)
Here c denotes the column-matrix with the following elements
ci = a2b3 - a3b2,c2 = a3bx - ai&3, c3 =axb2 -a2bx. (A.2.26)

Hence, it corresponds to the vector c which is


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
cf = —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 = 'Ea’ikXk (* = (A.2.30)
k=1
A.2 Operations on matrices 791

Premultiplying z with 1 x m row-matrix y' we obtain the scalar


nn
y'z = y ax = EE aikViXk = x'a'y, (A.2.31)
2=1 fc= 1
which is a bilinear form of the variables aq, £2, •••, x n and
1/1,1/2, •••, 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 5 3^2} •••5 %n

nn
(f(xi,x2, ...,xn) = -x'ax = x'a'x = EE Q'ikXiXfcj (A.2.32)
2=1 k = 1
formed by means of matrix a. This quadratic form is zero if
matrix a is skew-symmetric. Indeed, by eq. (6) we have

x'ax = x'a'x = —xax — 0.


For this reason we can omit the second term on the right hand
side of the following equation
x'ax — —x' (a + a') x + -x' (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
A = det a = \a\.
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.
mi x pi mi x p2 ... mi x ps
m2 x pim2 x p2 ... m2 x ps

mt x pimt x p2 ... mt x ps
Pi x mPi x n2 ... pi x nq
P2 x nip2 x n2 ... p2 x nq

Ps x ni Ps x n2 ... Ps x nq

(A.2.35)

where
Pi + ...+ps=p, mi + ... + mt = m, m+... + 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
Cu = (mk X Pi) (pi x m) + (mk x p2) (p2 x m) + . . + (mfc x ps) (ps x
m), (A.2.36)
which are mk xnj matrices. The order of matrix C is equal to (mi
+ ... + rnt) 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 matrices 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
llaa}]((aa)) + l[a(3}] [HI ((a/3)) + [[a0\] ((/?/?))
A P = (03a))
[[Pa]] (H) + [m\ ((/?«)) [\J3a]]((a0))+m]](m) (A.2.37)
Let us calculate the n x n submatrices of matrix AP. We have
[Ml ((««)) + [\a@\\ ((Pa))
n
^ {[as, ak\ (ak, ar) + [as, (3k\ (/3k, ar)} (s, r = 1,n)
k=1
A.2 Operations on matrices 793

Recalling the definition of Lagrange’s brackets (1.20) and


Poisson’s brackets
(1.26) we
n obtain
^{[as,afe] (ak,ar) + [as,/3k] ((3k,ar)}
k=l
Y' V' V' (dqm dPm dpm
- dq
™ \ (dak d(Xr dak dar
“ m~=\ tY\ V
das doLk da
s da
k) \ 9pt dpt dqt +
f dqrn dpm dpm dqm \f d(3k dar d(3k dar \
\ das d(3k das d(3k ) \ dqt dpt dpt dqt )
( dqm dar yr / dpm dak dpm
= EE d(3k \ +
( das dpt ^ V dak dqt d(3k dqt )
dpm dar ^ / dqrn dak dqm dj3k \ _
das dqt ^ \ dak dpt d(3k dpt )
dqm dar y^ / dpm dak dprn d[3k \ _
das dqt ^ dak dpt d/3k dpt )
dprndar y' ( 0qrn d(3k \
dots dpt ^ V dak dqt d(3k dqt ) '

Noticing that
Y' ( dprndotk dpm d(3k dp
^ V da dqt
k df3k dqt m = 0,
dqt
y- f dqrndak dqm df3k dq
\ dak dpt d(3k dpt
m
= 0,
y- / dpm dak dpm df3k dpt
V da dpt d(3k dpt
dpm _
dpt
k

y> / dqm dak dqm d(3k


\ da
mt 5
k dqt df3k dqt

we obtain
^^ &k] {p^ki Qr) “I” Pk\ {Pk^
^b*)}
k=1 1
nn dqm dar dp m da,
EX> das dqt das dpt
m=l £=1
y i'dar dqm + dar dpn dar
= —Sr
r'j V dqm das dpm das da.
An ^21 • • • An 1
A= A\2 A22 • • • An 2

A
A\n A2n • •• nnn

(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 = ^ ^ arkAsk,
k=1

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/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 |a|. Hence,
(aA)rs = 6rs\a\ or aA = \a\E. (A.3.2)

By analogy, we obtain
Aa = \a\E. (A.3.3)
2. Inverse of the matrix. Let us assume that a is a non-
singular matrix, i.e. \a\ 7^ 0 and denote
A=a~1. (A.3.4)
|a|
Then from eqs. (2) and (3) we obtain
aa~l = a~la — E. (A.3.5)
The square n x n matrix a~ is referred to as the inverse of
l

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~la~1 and postmultiplying them
by c~1 we obtain
b~1a~1cc~1 = b~1a~1abc~1
or, due to eq. (2.13),
b~1a~1E = b~1Ebc~1 = b~lbc~l = Ec~x.
Thus, if ab
= c then b~1a~1 = c~1.
This provides us with the formula for the inverse of a matrix
(ab)_1 = b~1a~1.
In particular
E-1=E=(aa~1) 1 = (a-1) V1,
and comparison with eq. (5) shows that (a-1) 1 = a, i.e. the
inverse of the inverse of a matrix is the matrix itself. Notice
that due to eqs. (5) and
(2.17)
\E\ = 1 = Mia"1!, i.e. la"1! = H"1 = A
(A.3.7)
\a\
E' = E= (aa-1)' = (a-1)'^.
796 Appendix A. Elements of the theory of matrices

Postmultiplying both sides of this equality by (a')-1


we obtain E{a)~X = (a-1)7 a' (a) 1 = (a-1)7 or
(a,)-1 = (a-1),> (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
^^askxk = bs (s = 1, ...,n) (A.3.9)
fc=l
can be written in the matrix form
ax = b (A.3.10)
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-1 we find

a xax = Ex = a 16, i.e. x = a 1b. (A.3.11)


Recalling eq. (4) we obtain an expanded form of the solution
^7L ft
Xk — i r ^ ^ Aksbk ^ ^ &skbk ($ 1? •••? ?
(A.3.12)
k=1 k=1
where the elements of the inverse e — oT1 are given by
e k — "j rAks fc — 1? •••? ^) •
(A.3.13)
s

\a\
4. Adjoint quadratic form. Let
^nn ^
<p(x i,x2,...,xn) = - EE dskXkXs =

(A.3.14)
5=1 k=1
be the quadratic form obtained by means of a non-singular
dp n
Vs = dxs ^Q'sk^k'j^ y = ax. (A.3.15)
k=1
The quadratic form is then transformed as follows
A.3 Inverse of the matrix 797

where due to eqs. (2.33) and (8)


e = (a”1)7 aa1 = E = (a~1)/ = a-1.
Hence
^x'ax = Iy'a~1y, (A.3.16)

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 tp (x±,... , x n) is denoted
as p'(yi, ... ,yn). Clearly
dp' ne
skyk (s = l,...,n),
dys Ek=1

where the elements esk of matrix a 1 are given by eq. (13).


Let us notice also the bilinear representation of the
quadratic form
n
-x'ax -x
2
11
y
2n X
sVs •)
k—1
where the latter equality, due to eq. (15) and (17), can be
written in either of two forms
<P= 2^2Xl dtp v' = liby^ (A-3-19)
Z
k=i dxs
k=1
expressing Euler’s theorem on homogeneous functions in the
case of quadratic forms.
5. Positive definite quadratic form. The quadratic form of n
variables
1, 1 n
if : -x ax = - ^ ^
2 2 a s kX s Xk
k=1
is called positive semi-definite if tp > 0 for all real values of the
variables. For example, the following forms
x\ + x\ + x\, (xi - X2)2 + (x2 — x3)2 + (x3 - xi)2
satisfy this condition. A positive semi-definite form which is
zero if and only if all its variables xi,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 x\ = X2 =
X3. If we consider the quadratic forms of three variables #i,
£2, x3 we conclude that the form
2, 2
functions

£1 = %1+ Ui X +
2 2 Ui 3 X 3 + ... FlLinX'
£2 = x2+ U 2 3 X 3 + ...
FU2
£3 = x3+ ... nX,
~{-U3 X'
n

€» = Xni

where Uik are expressed in terms of the coefficients ask of


form ip. With the help of the new variables we can write this
form as follows
<P = \ (Fifi + F2e2 + ... + Fnfn). (A.3.21)

Thus, the inequalities


an = Fi > 0, F2 > 0,..., Fn > 0
are the required necessary and sufficient conditions for the
positive definiteness of the form. These inequalities should be
expressed in terms of coefficients ask.
1 Ul2 Ul3 ... U\n
01 U2 • •• U2 n
u= 3 • • • ^3n
00 1

00 0 ... 1
Fi 0 . .. 0
F= 0 2
f . .. 0

0 0. .. Fn
an •••
&12 —1 0*1,71
o &22 ••• &2,n—1
21
Ol
n- , -1
1,2Q"n- ...
l &n—i,n—
an &12 = on.
&2 Its determinant is equal to unity. Introducing the diagonal matrix
1 022

we can put eq. (21) in the form


V= = \x'u'Fux
and thus
a = uf Fu. (A.3.23)
Let us designate the discriminant of form cp in the original
variables as An = |a|, and let An_i, An_2,... , Ai denote the
discriminants of forms <pn_2, • • • 5^1 which are obtained from
(p if we consequently set xn = 0, xn-i = xn = 0,... , x2 = x3 = ... =
xn_i =xn = 0.
It is clear that An_s are the principal diagonal minor
determinants of determinant An = \a\

(A.3.24)

It remains to notice that by eqs. (23) and (2.34)


An = \a\ = |F| = F1F2...Fn.
Putting £n = xn = 0 we obtain
An_i = FiF2...Fn-i.
800 Appendix A. Elements of the theory of matrices

For = £n-1 = 0, i.e. for xn~i = xn = 0 we have


An-2 = F1F2...Fn—2
and so on. The last equality in this chain will be

Ai=Fl
Then we find
Fi=A1; F2 = ^,..., Fn = ^±
f - ..rrif—
5n
~A
^n-1
and condition (22) takes the form
Ai >0, A2 > 0,An-! >0, An > 0. (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
that is, the signs alternate and an = A\ < 0.
Let us notice in passing that the matrix of the positive or
negative definite form is non-singular since An = \a\ ^ 0.
If An — \a\ = 0 whereas the all other A* > 0 for i = 1, 2,... , n —
1, then (p is a singular positive semi-definite form. It turns to
zero at ^ = £2 = • • • = fn-i = 0 and any = xn. For An = An-i = 0
and A* > 0 (£ = 1,2,... ,n — 2) we have Fn = Fn- 1 = 0 and the
positive semidefinite form is the sum of n — 2 squares £*,... ,
£^_2 multiplied by positive coefficients Fi,... , Fn~2, i.e. it is zero
at ^ = £2 = • • • = £n-2 = 0 and anY fn-i and etc.
5. Orthogonal matrices. A non-singular matrix a is called
orthogonal if the inverse a-1 of this matrix is equal to the
transpose a' of this matrix

For the orthogonal matrix ’


aa' = aa~ = E and a'a = a~ a = E.
l l

(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
(A.3.27)
matrix too. It
cd
X
II

II
"So
(A.4.1)
"C>

>-cT
(A.4.2)
II

II
1
cf

II
1

II
ab = ab', ba = ba'. (A.4.3)

Here c denotes the column-matrix (and correspondingly d the row-matrix) of


projections of the vector products c = a x b and a designates the skew-
symmetric 3x3 matrix (2.3).

The identities

a • (b x c) = b • (c x a) = c • (a x b)
can be cast as follows

a'bc = b'ca = dab,


whilst the identity
ax (b x c) = b (a • c) — c (a • b) = ba • c — b • ac
is written in the form
abc = ba' c — cab — (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
a2 = aa! — Ea'a.

(A.4.7)
Let us consider the linear functions of projections ai, a2, as of
vector a
Pn Pl2 Pl3
P2 P22 P2
1 P32 P3
P3 3
1 3

=P (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,
vector a being called the postfactor. The matrix notation of
this relationship has the form
c - Pa,
(A.4.11)
where P denotes the 3x3 matrix prescribed by the same table
(9) as that of tensor P, whereas c and a are 3x1 column-
matrices corresponding to vectors c and a. Transposing
relation (11) yields
which corresponds to multiplication of the transpose of tensor
P with the prefactor a
c = a P'. (A.4.13)
The matrix multiplication
— ^3^21 + —CI3P22 + —CL3P32 +
dP = &2^31 ^3^11 CL2P32 CL2P33 (A.4.14)
- CLlPzi — &3P12 — &3P13 “ ^1-
^33
a^Pw + aiP2i ttlP32 —
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 aPb in matrix
notation.
The matrix notation Pab corresponds to the following vector
(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 + At) -a(t) = ||aik (t + At) - aik (£)||.
A.5 Differentiation of a matrix 803

Multiplying the result by the scalar — we obtain


a(t + At) — a (t) &ik (j' "b At) Ojik
At (t)
Now we can calculate the limit for At —>At0, 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=||«ife||, (i = 1) •••) m\k = 1, ...,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' = 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
then the velocity vector is <71,...,
arguments determined
qn r =by
r(qthe column-matrix
1,...,q n;t) or x = x
• dx . dx . dx ,.^
V = X= — qx + ...+ — qn + —. (A.5.3)
d(J[ l (JQn dt
This equality can also be written in the form
dx . dx /A^,
V
x
~ dqq + ~dt' ( -5- )
where q denotes the column-matrix with the elements q s, s =
1,2, ...,n dx
whilst — is the 3 x n
matrix dq
dxi dx\ dx\
dqx dq2 dqn
dx2 dx2 dx 2 dx
dqi dq2 dqn dq (A.5.5)
dx3 dx
dx_3 3
dqi dq2 dqn
From eq. (4) it follows that
dv dx dx
(A.5.6)
dq dq dq ’
which is the matrix form of the following vectorial equalities
dv dr dr
Appendix B

Basics of tensor calculus

The analysis presented here is limited to basic knowledge. A


more comprehensive analysis can be found, for example, in
[47] and [51].

B.l General non-orthogonal coordinates


Let us consider three non-coplanar vectors ei,e 2,e3 which
form a vector basis. The vectors es 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
V — ei • (e2 X e3) = e2 • (e3 x ei) = e3 • (ex x e2)

(B.1.1)
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
e1 = -e2 x e3, e2 = -e3 x ei, e3 = -ei x e2
(B.1.2)
v v v
and form the dual vector basis. They are orthogonal to the
es • es = 1. (B.1.4)
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
e2 x e 3 = ^ (e3 x ei) x (ei x e2)
yZ
= ^ teie2 ' (e3 X ej) ~ e2ei ’ (e3 X el)] = “el
and the following expression
v\ = e1 • (e2 x e3) = -e1 • ei = -.

(B.1.5)
Then we arrive at the equalities
ei = — e2 x e3, e2 = —e3 x e1, e3 = — e1 x e2,
(B.1.6)
Vi V\ Vi
are of crucial importance. According to eqs. (3) and (4) we have

that is, \\gsk\\ is the unit matrix.


Let us prove that the symmetric matrices
9=\\9sk\\ and 5* = ||5sfe||
are the inverse of each other. Indeed,
3 3
3
1 9sk9km =
53 53e* ■ekek'eTn = e* ■ 5Zefcefe ■e™
k= k=1 k=1
1
3
= e • [em x (efc x ek) -f e e • efc]
s
k m

k=1
e x | >53
efcfex e"
e
= es + 9"

k=
1
It is easy to see
that
^3efc x e = i [ei x (e2 x e3) + e2 x (e3 x ei) + e3 x (ei x e2)] = 0,
k

3
k=1 (B.1.9)
B.l General non-orthogonal coordinates 807

and, hence,

YJ9sk9km=9T or gg* = E, g* = g~\ (B.l.10)


k=1
which is required. If follows from the above that
Gsk
9ks = "M"’ (B.l.11)

where Gsk is the algebraic adjunct of the element g sk of the


determinant |<7| of the matrix g. Alternatively, we can find that,
for example,

g12 = e1 • e2 = Uez2 x e3) • (e3 x ex)


v

and comparing with eq. (1) yields


= VW\- (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
9sk9km = 9?■ (B.l.13)
Using the values introduced by eq. (7) we can establish the
following relationships between the base vectors and dual
vectors
es = g ek, e = g ^ ,
sk
s sk
k
(B.l.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 el and e/, respectively. The result is
gls = 9sk9lk = 9sl, 9ls = 9sk9i =
9sU 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
^stq — x 6qr) and 6 ^ — e • (et x e ) . (B.l.15)
q
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
(B.1.16)
^=-
estq = VWl zStq =
If they correspond to an odd permutation of
1,2,3, then
(B.1.17)
estq = -VW\, estq = ~~j=.
Formulae (2) and (6) are now cast in the form
es x et = estqeg, es x = estqeq. (B.1.18)

B.2 Vectors using the non-orthogonal coordinates


Given a vector basis ei, e2, e3, we can describe an arbitrary
vector a twofold. We can represent it in the original basis
a = ases

(B.2.1)
or we can prescribe the three scalar products
as= aes (s = 1,2, 3).

as |es| = a ^/gZ
(B.2.2) s

(no summation over s) are equal to the edges of the


parallelepiped constructed on the base vectors. The vector a
is the diagonal of the parallelepiped and the line segments
QJ S CLS
\/ 9ss
|es|

are the projections of a onto the base vectors.


Using eqs. (1), (2) and (1.14) we obtain the equations
relating the covariant and contravariant components
cls — 9sk& 5 & = a • e — Q
(B.2.3)

a = akek = akgkses = ases, (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
as = gskak and es = gskek
demonstrates that the relationships between the covariant
and contravariant 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
as = gskak and es = gskek
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 , e'2, is assumed as a new
basis.
The difference between the covariant and contravariant
components disappears for an orthogonal Cartesian
coordinate system with |es| = 1.
The scalar product of two vectors can be set in any of three
forms
ab=a
s
bs = gskasbk = gskasbk. (B.2.5)

In particular,
a2 = asas = gskasak = gskasak. (B.2.6)
or
c = c et
l
= asbkes x efc = asbkesktet.
Therefore, the covariant and contravariant components of the
vector product are equal to

c* = e
skt
Ct = esktasbk, asbk. (B.2.7)
For instance,
Cl = V\g\ (a2&3 - a3fr2) >V\9\
c1 = —(a &3 - a3b2)
2
(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

esek, esek, esek, esek,

which gives rise to the four dyadic representations


P = P ee =P ee =P ee =P ee
sk
s k sk
s k s
k s
k
s
k s
k (B.3.1)
by means of the contravariant P sk, covariant Psk and mixed
Psk,Psk components.
Using eq. (1.14) it is easy to obtain the relationships
between the above components. Postmultiplying eq. (1) by e m
and el we arrive at the equations

plm = gSlgkmpsk = gkmplk = gSlprn (B.3.2)


By analogy, we obtain

Plm = gsl9kmPSk = gslPSm = flfen.P/*, (B.3.3)

Plm = glSPsm = 9kmPlk = (B.3.4)

Pm = gslPms = 9kmPkl = 9ms9klPSk- (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 P sk = Pks as well as for the mixed
components Psk = Pks. Hence we can adopt the following
notation Psk = Pks — Pk 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 gsk, contravariant gsk and mixed
gk components are defined in terms of the base and dual
vectors by means of eq. (1.7).
c = Pa (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
c = c e = P e e ■ a„
s
s
s
k s
k = Pskakes,
that is,
cs = P'kak
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
= —Pfcs, then Psk = — Pks and Psk = —Pks- A skew-symmetric tensor
is given by three components. Introducing the Levi-Civita
symbols, eq. (2.1), we can assume that
psk = psk = -tlskL0l. (B.3.7)

Postmultiplying the skew-symmetric tensor P by a we have


c=Pa
or

cs = Pskak =-elskakujh cs = Pskak = -elskakto\ (B.3.8)

which is equivalent to the following extended form


c1 = —1= (w2a3 - W3a2), Cl = y/\g\ (w2a3 - u>3a2)
V\9\
and so on. We can arrive at the same equation if we enter
vector u) and calculate the vector product lo x a. Hence,
adopting notation (7) we obtain that for the skew-symmetric
tensor
P • a = (jj x a.
(B.3.9)

B.4 Curvilinear coordinates


The position of a point in three-dimensional space is
prescribed by three quantities , <?2, <?3, referred to as the
curvilinear or generalised coordinates. This means that the position
vector r should be considered as vectorial function of these
quantities
r = rtf,qW). (B.4.1)
dx dx dx
2
d,qi dq dq3
dy dy dy
dq1 dq2 dq3
dz dz dz
1
dq dq2 dq3

within the range of the variables q1, 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
dr (s = 1,2,3). (B.4.4)
rs dq
s

These vectors are non-coplanar as the value


V = n • (r2 x r3) (B.4.5)
is equal to the Jacobian J and, thus, is not equal to zero.
Let the vectors rs and rs 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 aby
dr = radq , (B.4.6)

and, thus, the square of the distance between these points is


equal to
ds2 = dr • dr = rs • rkdqsdqk = gskdqsdq
k
.

(B.4.7)
Hence, the covariant components of the fundamental tensor
can be expressed as the coefficients of the quadratic form ds2.
These determine the metric of the chosen system of
curvilinear coordinates in the vicinity of the point under
B.4 Curvilinear coordinates 813

Let us begin by calculating the following vectors


d2r _ dr __ d*s _ k
Tsk
^dqk dqkdqs ~ dqs ~ ^
They can be represented in terms of the base vectors as
follows

rsfc =
{5fc}rm‘ (BA9)

The coefficients designated by the braces are referred to as


Christoff el’s symbols of second kind. Due to eq. (8), they are symmetric
with respect to the lower indices

(B.4.10)

If follows from eq. (9) that

rst * rt = 9mt (B.4.11)

and then, by virtue of eq. (1.13), we have


= gtlrsk -rt. (B.4.12)

The scalar product on the right hand side is denoted as


• rt = [s, k; t] = [k, s; t]. (B.4.13)
These values are referred to as Christoff el’s symbols of first kind. The
following notation

[a,M = rsM) m = r‘fe


is often used for Christoffel’s symbols. By virtue of eq. (13) we
have

~HqiVs'Vk = =
+
By means of the circular permutation we also obtain
-jfff = [k, s; t\ + [f, s; k], = [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
1 / dgst dgkt _ (B.4.14)
[s, k; t] dgsk
\dqk d(f
2
With the help of eqs. (12) and (13) we
have
The sequence of eq.(15) is the following (B.4.15)
equalities

[S,M] =
(B.4.16)
A simple way to obtain the derivatives

dr* (B.4.17)
r14 = dq
C s
is as follows. Due to eq. (9) we have
a
cir9‘=0 = ~Sq’r‘' * ” r < '
r
“ + 1 lr”* ’r
r •— -- ■*

or

• ri =- (B.4.18)
si
This means that the values on the right hand side can be
formally treated
as the covariant components of the vectors r*. Thus

(B.4.1
9)

Let us notice that Christoffel’s symbols are zero if and only


if gtm are

B.5 Covariant differentiation


In mechanics and mathematical physics, the invariant
quantities are of interest. 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 (velocity, 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 quantities 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 basis 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 forthcoming 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.
<9a
dqs ivy

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
da dak k da171
7T~ = /c ~7\ T +
a
r
(B.5.1)
dqs
Using formulae (4.19) we also obtain that

hence,

(B.5.3)

where the expressions

(B.5.4)
816 Appendix B. Basics of tensor calculus

are called the covariant derivatives of the contravariant and


covariant components 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
—P = rkrtVsPkt = rVvsPfct = rfcr‘VsP],

(B.5.5)

where the covariant derivatives of the quantities Pkt, Pkt, Pit


are

(B.5.6)

These formulae enable the covariant derivatives of the akbt of


components
a dyadic to be calculated. It is easy to see that the rule
Vs&kbt — sbt T s&k
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,
VsPktQml = PktVsQml + QmNsPkt-
The covariant derivative of a scalar is given by
_ d(p

Due to eq. (6), in the case of the scalar product of two


vectors we have
Vsakbk = -J^akbk + ( k la” oqs
ism) bk
~ ^ ?k}akbm =dik~sakbk’

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
dgkt
^ sQkt — dqs [s, k\ t\ — [s, t; k\ = 0.
(B.5.7)
The mixed components are either equal to zero or unity, thus,

Calculating the covariant derivative by means of eq. (6) we


obtain
Vs9i=
-{Tk}9™+{sL}9™=■{!}+{1}=°-
It remains to consider the contravariant components. Using
relationship
(1.13) we obtain, by virtue of eq. (7), that

Vsg{ = 0 = Vsgkmgmt = gkmVsgmt.


Given s and £, we obtain a homogeneous system of linear
equations with the non-zero determinant \g\, hence
Vsgmt = 0.
Summarising we see that
v,9kt= 0, Vsgkt = o, Vsgi= 0.
(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 Vs. We can see from formulae (5) that
rfcr VsPkt = TkTtVsgkmgtlPml=gkmTkgtlTtVsPml
t

= rmr'VAi=rVvsPt(.
Let us consider now the tensor
a da
r ---- ir-------1-
9a 9 da o z9a
r ----1- r' --
3 (B.5.9)
dq
s
dq 1
dq dq ’
2+ 3

which is a sum of three dyadics. Formulae (3) yield the dyadic


representation of the tensor
818 Appendix B. Basics of tensor calculus

The values Vsam and Vsam are the coefficients of the dyadic
representation of tensor (9). Thus, they can be considered as
the components of the tensor, namely, V sam are the mixed
components (covariant and contravari- ant with respect to s
and m respectively), whereas Vsam are the covariant
components with respect to both indices.
The partial derivatives of both a1 and a* 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 example. 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 p = rz (r is constant). Let two
orthogonal axes Ox1 and Ox2 lie in the plane of the plate, then
any point in space is typified by the three coordinates x^x^x 3 =
z which can be understood as g ,# ,# . Equation (4.2) ex-
1 2 3

pressing the Cartesian coordinates x,y,z in the fixed axes in


terms of the generalised coordinates takes the form
x = x1 cosrx — x2 sinrx , y = x1 sinrx + x2 cosrx , z = x3.
3 3 3 3

(B.6.1)

Thus, the projections of the base vectors of the covariant


basis on these axes are given by
ri : cos tx3 sinrx3 0
r2
r :
: — sin rx3 cos rx3 0 .
3 ~ry tx 1
The covariant components of the metric tensor are equal to

9n = 1, 9i2 = 0, 513 = -TX2,


522 = 1, 523 = 12 22 (B.6.2)
533 = 1 + T (rc ) +(x )
2
TX1, .

Now we have
10
=1 (B.6.3)
—TX2 TX 1+T2 (x1)2 + (x2)2

and

(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] = -rcV2, [3,3; 2]= -x2t2 1
[2,3; 1] = [3,2; 1] = —T , [1,3; 2]= [3,1; 2] = r, \
(B.6.5)
[1,3; 3] = [3,1; 3] = x‘r2,[2,3; 3] = [3,2; 3] = -
x2r2. J
It remains to construct expressions for Christoffel’s symbols

whilst the other symbols are zero. The calculation can be


simplified by direct determination of the derivatives of the
base vectors. This3 yields 3
iq = ii cosrx + i2 sinrx , r2 = — ii sinrx3 + i2 cosrx3,
ii = ri cosrx3 — r2 sinrx3, i2 = ri sinrx3 + r2 cosrx3, r3
= r (-iiy + i2x) = r (x1!^ - x2ri) . (2.6.7)
For this reason
i'll = ri2 = r22 = 0, 1*13 = t (-ii sinrx3 + i2 cosrx3) = rr2,
r23 = -ti*i, r33 = r (x1r23 - x2ri3) = -r2 (x1!*! + x2r2) .
(B.6.8)

Now, using the formulae in eq. (4.9) we arrive at the same


expressions (6) for ChristoffePs 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 q1, q2, respectively, we can cast
the vectorial equation of the surface in the form

p = p(<?\g2) (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 coordinate lines of the surface. Then we can determine an
infinitesimally small displacement on the surface and the
square of its value as follows

dp = padqa, \dp\2 = (da)2 = pa ■ pf}dqadq0 = aal3dqadq0, (B.7.3)

where
aa0 = Pa Pf3 (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
11 &22 12
a = T T , a = - T T , a =-r-r,
a
12 22 all (B.7.5)
a a a

where \a\ is the determinant


aiiai2
a= — &11&22 “ &12> (B.7.6)
dl2 &22
and the following equality holds

h ! 44: (B.7.7)

The dual vectors


Pa = aa0P/3 (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
Pa P0 = aa0 (B.7.9) holds, that is p1 and p2 are perpendicular to px
and p2- Let us notice that
\PI x p2I
2
= P\PI - (PI ■ P2)2 = M,
(B.7.10)

i.e. the element of the surface area is equal to da = ^/\a\dq1dq2.


Any vector c on the surface (i.e. the vector lying in the plane
tangent to the surface) can be prescribed by the contravariant
ca and covariant ca components
C = Capa = C(% pa, (B.7.11)
B.7 Formulae of the theory of surfaces 821

where
cQ = c ■ pa, ca = c ■ pa. (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
1 surface is defined as follows
m = Pi P2
x
Pi X p , 2
(B.7.13)
\Pl X
hence, P2I

Pa ■ m = 0. (B.7.14)
taking derivative with respect to q13 we obtain

Pa{3 ' m — ~Pa ' m/3 — ba0 — b0a — pa0 ■ (p X p ) .


1 2 (B.7.15)
vM
The quadratic form of differentials dqa of the Gaussian
coordinates
I = (daf = aa0dqadq0, II = ba0dqadq0 (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 bap of the
second quadratic form are the covariant components of the
tensor, the mixeda components are
b} = <r ba0 = —a,apct ■ m0 = -p7 • (B.7.17)

and the contravariant components are given by


b60 = aSaa^bS7.

(B.7.18)
We proceed now to obtaining formulae for the second
derivatives paj3 of the radius-vector p. As indicated above, the
vector paj3 is not a surface vector and, thus, its representations
should contain a component along the normal to the surface.
Assuming
Pal3 = + ba0m

(B.7.19)
Pa0 ■ PS ~
(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 / daaS daps daa0 (B.7.21)
*>«, ■ P, = Kft<5] = 2 + ajr -
It follows from eqs. (20) and (7) that

>■ (B.7.22)

The values
[a, <5]

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
mQ • m = 0, (B.7.23)
which indicates that the derivative of the unit vector is
orthogonal to this vector. For this reason, m a, being a surface
vector, can be represented in terms of the vectors p@ or p^.
Accounting for eqs. (12), (15) and (17) we have
a = -bapp
0
m
= ~b0Pp- (B.7.24)

Let us construct the formulae for differentiation of the dual


dp vectors, i.e.
dqP
Repeating the derivation resulting in formula (4.18) we find that
Then we have
Pp m = 'm = = bp,
(B.7.26)
as the other components of the vector obtained by
differentiating a?*1 p1 are orthogonal to m. The consequence of
eqs. (25) and (26) is the required relationship
^ = -{/3a7}^ + ^m-
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 Q1 = Qcoordinates
1
(t), q2 = q2 (t), 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 q1 and q2 are given by eq. (!)•
In what follows, the arc length a of the line in question is
taken as an independent variable. The unit vector r of the
tangent is equal to dp da
T=
Ta=P (B.8.2)

Denoting the curvature of the line and the unit vector of the principal
normal by and n, respectively, we obtain with the help of
k

Frenet’s first formula

d 2p dr d2qa dqa dqP


= kn : = Pa()
da2 v“ da Pa ~d^ da '

,kn = p
da22qa
(d a 1 dqP dq1 . dqa dq13
An^ +
+ bot/3—] —m- (B.8.3)
[/?7j da da da da
The vector of curvature has a component along the normal to
kn

the surface
r, t dqa dqd bafjdqadqd
mk = mkn ■ m = mbap— — = m------ g, 3 (B.8.4)
da da a^dq^dq
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 r. Due to Meusnier’s theorem, the curvature of
the normal section of the surface is given by

=
k kn • m = k cos 6.
k* = kn—km = pQ (B.8.6)
(B.8.5)
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
d2qa ( a 1 dqP dq1 0 (a = 1,2). (B.8.7)
dcr2 ^ \/?7j da dcr
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
their
of twoderivatives
differential
da equations (7) for given initial values of
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 = zbm and k = ±fc, that is, k* = 0.
The geodesic curvature of the surface line is the value of the vector of
the geodesic curvature, i.e.
|k *\ = Jaa0k*<*k*fi. (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 derivation. The covariant components can be
represented in the form
dqP dqa
da da (B.8.9)
d

where <3> denotes the following quadratic form


1 dq dq
a 13
,
*2= (B.8.10)

Using formulae (7.27) we obtain


. dn dr a dq
13
kn = — +a rapa0-
—rap“ = p
d da
0
a
adra dq ( la,
p +b
<>m
dTa _ ( 7 ) d<f_ ° °^
+T b m
=P da l/3aj da 7

or

d<f> dqP dq6 dqP dq1


kn = pa
da dq \ da daai6\fda
a
d
da
B.9 Covariant derivative of a vector on the surface 825

Recalling the

relationship and

noticing that
\ _ 1 /daps daas dapa \ dq& dq6
[p, a, \ ^^ 2 \ dq dqP
a
dq6 ) da da

d< + mb. adq


a
d d<S> dqP
kn = p p —— —. (B.8.11)
a

$> da da
dq“
Expressions in the brackets are the covariant components of
the vector of the geodesic curvature and are zero along the
geodesic lines. Hence, another form of the differential equations of the
geodesic lines is available
d<$> d<f> (B.8.12)
- — =0 (a = 1 , 2 ).v
da d dq a
dqa
da
We notice also that vector r and the vector of the geodesic
curvature are orthogonal since
r • k* = r • (kn — Amm • n) = 0. (B.8.13)

B.9 Covariant derivative of a vector on the surface

Let us construct the expression for increment dc in the vector c


prescribed on the surface by its contravariant components ca.
Applying the formula of differentiation (7.19) we obtain

dc —da? — da13 d cap


QqP q q d(f Pa
tfpo (^8 + {7^}c7) +

The term
dq^cabapm (B.9.2)
826 Appendix B. Basics of tensor calculus

is the component of dc along the normal to the surface,


whereas the term
d* c = dqppaWgca
fa1
dca
V0ca =
dqP + \/3jy (B.9.4)

denotes the covariant derivative of the contravariant quantities


cQ Along a line L on the surface we have
aa/\ . a dqp dca dc° dqP
c =c w, ^ = =

and the expression for d* c takes the form

' - ipp° (t? +


dc
*r) =p• (*“+ {^d}^ (B.9.5)
The vector is said to have translated parallel along the line L if
d* c = 0 i.e. V fiCa = 0.
(B.9.6)
In particular, when c is the vector of the tangent to the line L,
then, due to eq. (8.6),
a dqa
(d?q f a a (B.9.7)
C
=~i,7’ d T = dPp‘ {+ da
With this in view, let us consider the change in the scalar
product c • r
along L. Taking into account that c and r are the surface
c m = 0, r • m = 0
and, therefore,
d(c • r) = c • dr + r • dc = c • d*r + r • d*c. (B.9.8)
It follows that under a parallel translation of c along the line L

we have
d{ c • r) = c • k*d<r, (B.9.9)
and, specifically, when L is a geodesic line, then
d{ c • r) = 0.
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
(B.9.11)

where the values in the parentheses are the covariant


derivative of the covariant quantities ca. Repeating the
reasoning of the above transformation we obtain

(B.9.12)

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 geometry 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
second 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 introducing2 the vector of the normal to the
p{q\q ) = v{q\q\ql).
Then

and
1
(B.9.14)
828 Appendix B. Basics of tensor calculus

Here the relationship (1.2) is used. Now we have


d1 1 dr3
/“aa /“aa fi„B (B.9.15)

Noticing that

we arrive at the equality

(B.9.16)

Here
i3 = 9o K P\ 7]0 + 9o3 K ft 3)o (B.9.17)
\a0 0
and, in order to calculate [a,/?;3] 0, we need the derivative of
gap with respect to q . Thus, determination of the coefficients bap
3

implies knowledge of the components of the metric tensor ga3


at q3 = q$ and the components of 9a(3 up to terms of the first
order in q3 — <2$
(B.9.18)

We have to ” leave” the surface as it is not possible to


determine the second quadratic form while ”staying” on it.

B.10 Orthogonal curvilinear coordinates


When the base vectors r i , r 2, r 3 are orthogonal, the non-
diagonal components gsk of the metric tensor are zero.
Denoting the diagonal components gss by h2s we have

(B.10.1)

where hs are referred to as Lame 7s coefficients. Then we


have
(B.10.2)
M = hjhlhl,
B.10 Orthogonal curvilinear coordinates 829

and thus
9 ** = i r* = I|.
(B.10.3)
The expression for the vector a can be cast in either of the
following forms
a = asrs = asrs = j^Ts-

(B.10.4)
Therefore
of
_ a(s) = ashs (B.10.6)
h~s
with no summation over s. ’
The square of the element of the arc length has the form
(ds)2 = hi (dq1)2 + hi (dq2)2 + h\ (dq3)2 ,
(B.10.7)
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 f s 1 dlnhs ' dq ’
k
[s, s; k\ — hs , hs dhs
[s, k;s] = [k, s; s] = , hi dqk
’ d In hs (B.10.8)
dhs dq s
‘ ,
[s, s; s] = h: s
dq

As an example, let us consider the spherical coordinates
R,d,p, where R denotes the radius of a sphere on which the
point lies, d is the angle along the meridian from the north pole
(0 < d < 7r) and p is the azimuth of the meridian from the plane
Ozx to the plane Oyz (0 < p < 2TT) .
The Lame’s coefficients are equal to
hR = 1, h# = R, h^ — Rsind, (B.10.9)

as the lengths of the arc element of the radius, meridian and


the parallel circle are dR, Rdd, R sin ddp, respectively. The base
vectors are given by
Lr = i r, = Ri$, r^p = R sin dip, (B.10.10)
830 Appendix B. Basics of tensor calculus

where the unit base vectors IR, i#, have positive directions of
the radius, the tangent to the meridian and the tangent to the
parallel circle respectively. Now we have

}rK + {rrY° + {/flK = °'


}r*+{jt}r' + {™}r*’ = :Sr<=i'’
' I Rifj * \Ry>
r
" = {»} r B + W I r ' + = rv = ivsin??,
~RrR'
| R\ [ 1? ,
=
Up}TR + Up >r»+ cot $ = R cos
[ . {;} h w)
+
\ R\
rw = < >rR + < >r,9 + < ‘ sin •& (Ryr sin i? + cos
>r, $)
(B.10.11)
There is no need to repeat the derivations carried out for the
orthogonal
Gaussian coordinates on the surface. The unit base vectors —
p\, — On, m
hi h2
hi h2 1
Pi = Y2P2 X m’ P2 -Pi x m, m = h\ hih2 Pi X P2-

It follows that
bi2 = -p2 • m i = ^ ( p 1x m ) ’ mi.
h\
Let us consider the three vectors px, m and m* = m + midq1. The
latter is the vector of the normal to the surface at the point (q1 +
dq1, g ) which is infinitesimally close to the point [q1, q2) under
2

consideration. The vector product

(px x m) • m* = (p± x m) • mxdq1,


which is proportional to 612, is the volume of the parallelepiped
constructed on the vectors px, 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, 612 = 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.10 Orthogonal curvilinear coordinates 831

Turning to formulae (9.17) and assuming that the surface in


question corresponds to the value q3 = of the orthogonal
system, we obtain

So7 = 0, V^=(/^jy [1-2;31=0’ [a,a;3] = -ha ^

with no summation over s. Hence,

The latter equality shows that the surfaces of the orthogonal


system intersect on their curvature lines. Assuming q1 = d,q2 —
tp,q3 = R for the spherical surface of the radius i?o, we obtain

= —Ro, = -R0 sin2d

and the normal curvature, due to eq. (8.4), is given by


£_ ((I'd)2 + bw (dtp)2 __________1_
hi (dd)2 + h2 (dp)2 Ro
In accordance with eq. (11), there are only two non-zero
Christoffel’s symbols
< \ = cot d, \ ^ \ = - sin d cos d,
[dtp) [tptp J

and the covariant components of the vector of the geodesical


curvature are as follows

k*v
d2d dt 2sin d cos d2tp — dtp dd . 2-~
cot d. da
da 2
p k*^ =
lb 2 da
da $,
For example, along the meridian

da = Rodti, ||=0, ^ = 0, i.e. k*® = ft** = 0,


da1 da
i.e. the meridian is a geodesic line. For a parallel circle
1
da = RQ sin ddtp, k*® = — cot d, k^ — 0
R0
and, due to eqs. (8.8) and (8.3)
r= C O t ^ k= h 2 + k*2 = 1

RQ RQ s i n d ’
832 Appendix B. Basics of tensor calculus
As a second example, let us consider the surface of a right
circular cone
d = $Q. Then, by virtue of eq. (12), we have

b<p<p = -R sin $0 cos $0, bRR = 0

and the normal curvature of the line on the cone is


cot $o 1
k 2 • dR\
R __________
1 + R2 sin2 $o \dp)
The non-zero Christoffel’s symbols are

PI Rip -
= —R sin2 $o-
\WJ R'
The contravariant components of the vector of the geodesic curvature are
d2(p 2 dp dR dp
k*» = d^-R
2
k«P =
R sin2^ ^
daz Rdada d 2c 1
da
and the geodesic curvature of the parallel circle (da = Rsiw&odp) is
R~x.
Let us also study the case of a surface of revolution. The
position of a point on the surface is given by the arc q1 = s along
the meridian and the azimuth q2 = p which is the angle between
the meridian plane and the assumed plane q2 = 0. Lame’s
coefficients are given by
where r (5) 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
g iz- r r ' ( ^ y = 0 , p§+2r-^±=o,
da \da) daz r da da

where a is the arc along the geodesic line. The first two
integrals are as follows
di
Pr2_r (ds\2_r C1 TaT "Cl’ l^J ~ °2
~ 2^2’

and the solution reduces to the two quadratures.


B.ll Finite-dimensional Euclidean space 833

B.ll 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 x i , . . . , xm. Let us agree to understand
these num-
bers as the coordinates of a point in a Cartesian orthogonal
system of axes,
then the set ( x i , . . . , x m ) corresponds to a space of the
dimension m. If
the distance M N between two points of this space M (x 1, . . . , x m )
and
N ( y u . . . ,V m ) is given by

M N = y j ( x \ - y i f + ... + (xm - ym) ,


2
( B. l l . 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 E% is valid in Em- For
example, the
coordinate transformation with the same origin O is carried out
by means
of the following formula
m
x' — ax or xfs = ^dsfcXfc, (B.ll.2)
k=1
where a is an orthogonal m x m matrix describing the
transformation of
one coordinate system to another (rotated) system. Here xf
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 Ox*, whereas the values (B.ll.3)ast
for a taken
s determine the projections of the unit base vector i' of the
834 Appendix B. Basics of tensor calculus

The scalar product of two vectors is defined as follows


m
a - b= ^ o s 6 s .

(B.11.4)
5=1
The concept of the tensor of second rank is introduced in Em by
analogy with E3. It is a set of m2 values Pst which is transformed
due to the rule
m
bs = Y,Pstat, (B.11.5)
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
mm
p
P=X)S “i»i‘- (B.11.6)
5=1 t=1
An example of the tensor is the dyadic product of the
vectors
mm
ab = ££» sbtisit• (B.11.7)
5=11=1
A curve in Em is given by the parametric equations
xs = xs (t) and r = r (t). (B.11.8)
For instance, the arc a along the curve can be assumed as the
m
da = X (dxs)2 (B.11.9)
N
The vector
dr dxi
r = — with the projections = ——
da da
determines the unit vector of the tangent to the curve. Its
derivative defines the vector referred to as curvature vector
£;n, which has the direction of n
d2 r d2xt
— or kn, = -^, ( B . l l. l l )

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.ll Finite-dimensional Euclidean space 835

It is clear that the non-orthogonal coordinate systems with


the base vectors es can also be introduced. They determine
the tensor with the covariant components
gsk=es-es (s, k = 1,... , m).

(B.11.12)
Its contravariant components gsk are determined by
relationships (1.11) as elements of the inverse for matrix |gsfc||-
By virtue of the properties of the inverse for a matrix, the
mixed components
9sk9kt=9ts (B.ll.13)
are given by eq. (1.8). Then, eq. (1.14) yields the vectors of
the dual basis
es = 9 et,
st

so that we obtain, with the help of eqs. (1.12) and (1.13), that
e s - e t = g ? , e ■ e* = g g*glq = g g\ = g .
s sl sl st
(B.11.14)
The vector a can be prescribed by means of one of the
a = ases = ases, as = a - e s , a s = a - e s , (B.ll.15)
i.e. by means of contravariant as or covariant as components. In
contrast to Secs. B.1-B.3 the indices, the dummy ones
included, take values 1 , . . . , ra rather than 1,2,3.
The curvilinear coordinates q1 , . . . , q771 can also be introduced
in Em. The content of Secs. 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
('ds)2 = gskdqsdqk (B.ll.16)
can always be transformed to the sum of squares
( d s ) 2 = ( d x i) + ... + (d x m) .
2 2
(B.ll.17)
To this end, it is sufficient to return to the Cartesian
coordinates x s by
means of the equations which are the inverse to the following
ones
x s = x s ( q 1 , . . . ,q ) or r = r ( q 1 , . . . , q m ) .
m
(B.ll.18)
This inversion is always possible when the determinant
dx\ dqd x7\
\A9\ = W ' 1
(B.ll.19)
8xr
dq
1
” * dqdx m r

is non-zero.
836 Appendix B. Basics of tensor calculus

B.12 Riemannian space of dimension n


The relationship
xs = xs ( q 1 , . . . , q n ) ( s = 1 , . . . , n ,... , m ) (B.12.1)

or in vectorial form
r
= r (g1,... ,qn) =p(q1,... ,qn) ,

(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 E$ is considered as
space i?2- Let the Greek letters denote the numbers 1 , . . . , n
and the indices whose values are greater than n be
designated by n + L n + q etc., the Latin letters t, q etc. taking
dr = ——adqa = dqa (B.12.3)
dq dqa vJ

separates the set belonging to Rn from the set of infinitesimally


small vectors at the point M (x 1, . . . , x m ). The vectors
dp
Pa dqa (B.13.4)

form a coordinate basis of the space R n in the vicinity of point


M. As R n belongs to E m , the square of the arc element is given
by
( d s ) 2 = ( d x 1)2 + ... + (dXm)2 = -7^dq •
a
= aapdq dq .
a p

(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 ca
(dxa) ,
a=l
a few of the constants ca being negative. In the latter equation
aa(3 — Pa' P(3 (B.12.6)
denotes the covariant components of the metric tensor. Its
contravariant components aa(3 are determined as elements of
the inverse to the matrix |aa/tf||. The inverse exists as the
quadratic form (5) of the differentials dqa is positive definite.
Given a0^, we can construct the dual basis
Pa = aa(3p0 (B.12.7)
B.12 Riemannian space of dimension n 837

at point M. A vector c is said to belong to the Riemannian space Rn, if it


can be expressed in terms of the vectors pa or the dual vectors
pa

c = capa=capa. (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 qs to m variables
x±,... ,xm (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 since we can
introduce (not uniquely) such linear functions pa of the
differential dq@
tpdqP = dpa (B.12.9)

that the expression for (ds) takes the form


2

(,dsf = (dp1)2 + ... + (dpnf . (B.12.10)


System (9) is not integrable in the general case as p are a

differentials of the quasi-coordinates. Hence, only a local


expression for (ds) in form (10) is possible since relationships
2

(9) are integrable for constant values of fg, i.e. for fixed values
of q1,... , 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
(i.e. Em). As illustrated in Secs. B.7-B.9 by an example of the
manifold R2 in E3, it is necessary to differentiate the internal
properties of R2 due to the given tensor from the relationship
a
dc V7 _ d° a
W = ^8 +
a V/3Cq
dq* (B.12.11)
dqP \h]a

where index a indicates that Christoffel’s symbols are


calculated in the metric of Rn
lnc,6 \dap6 dap^y
2 [ dq i dqP (B.12.12)

Using eqs. (11), (9.5) and (9.12) we can construct the


expressions for the increment d*c of vector c in the space Rn
when it moves to an infinitesimally close point of the line qa =
qa (a). However it is necessary to remember that the increment
dc of this vector in Em differs from d*c in the components which
are normal to the base vectors pa 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 translation of vectors introduced in Secs. B.8 and B.9
for R2 remain valid in the space Rn as well.
If one of the base vectors, say p^, is assumed to replace the
vector c in eq. (9.5), then the covariant components ca should
be replaced by
/ 0 (a ^ 6),
\ 1 {a = 6).

This yields the


relationship
(^} V = dql3Vfjpg.
d
(B.12.13)
*Ps = Poc
The introduced vectors
V/5P
* = {^} Pa =
(B.12.14)

can be referred to as the covariant derivatives of the base


vectors p in the space Rn. Due to the symmetry of Christoffel’s
6

V/3P6 = V6Pp. (B.12.15)


Relationships (14) replace equations (4.9) in the space Em.
Any calculation 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
V/3C = pa (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 Riemannian space Rn determined by the square
of the linear element (or the covariant components of the
metric tensor aap) reduces to the search of vector p^1,... ,<?n)
in Em from equalities (12.6) with prescribed right
839 En
B.13 Riemannian subspace Rn in the Euclidean space

hand sides. The number of equations is \n (n + 1) and the


number of unknown variables is m which is the number of
projections of vector p on the coordinate axes in Em, see eq.
(12.1). If
rn = (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 E%
whilst in general R% is embedded only in Eq. 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 Rn.
The curvilinear coordinates q1,... ,gn,... , qn+P are introduced
such that the radius-vector p (g1,... , qn) of the points in En+p
belonging to Rn corresponds to the fixed values of the
coordinates gn+1,... , qn+p, i.e.
The base vectors in En+p are denoted as
Ta=
dr' rn+s=d^n (<x=h---,n-, s = l,... ,p). (B.13.2)

The system of base vectors in Rn is given by

Pa = (^)o = (ra)°,
(B.13.3)
where zero denotes that the value defined in En+P is considered
ga0=ra-r0, [a, (3; 7] 1 f dgai3 (B.13.4)
dffq7 dgp7 dq'i
in Rn are respectively equal to
1 / daoc
(daft)0 QJctf3') [o^,da/3,
a7
'T]0 da 7]a-p
( 37 dq'y
(B.13.
Having the covariant components of the metric tensor in5)En+P,
i.e. the matrix with the elements
9ot(3 — • r^, ga^n+t — ra • rn_|_^, $n+s,n+£ — rn_|_s • rn_|
_^,
we find the contravariant components which are the elements
of the inverse for this matrix
at/3 a,n+t n+s,n+£
y5y1y
840 Appendix B. Basics of tensor calculus

With the help of these elements we find the vectors of the


dual basis
r“ = g<*% + ga'n+trn+t, vn+s = gn+s’ara + gn+s'n+t vn+u (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 gf with aaP or rg with aaf3pp =
pa is a grave error. Fully analogous, the value

{;7}o = r,Ha + 9o’n+t [ftr,n + i]0 ± ,


which is Christoffel’s symbol of second kind defined in the
metric of En+P but at points of Rn, is not equal to these symbols
calculated in the metric of Rn.
It is known that the vectors TQ+t are orthogonal to ra, hence,
at points of En+P belonging to Rn we have
‘ ro+t = 0,
Pa (B.13.7)
i.e. are p vectors in space En+Pwhich are orthogonal to all
vectors pa,
and thus orthogonal to Rn. It is obvious that any vector which is
linearly expressed in terms of rQ+t is orthogonal to any pa. We
can prove the converse statement, namely, if m • pa = 0 for a =
1,... , n then vector m is a linear form of vectors rQ+t.
Taking into account the above, let us assume the vectors
ra, rn+s (B.13.8)
are the base vectors of the space En+P. Let us express the
second derivatives rap of the vector r in terms of these base
vectors
= f>7 raP + qa \n+trn+t. (B.13.9)
0

The coefficients are not Christoffel’s symbols of second kind.


Indeed, the scalar product of eq. (9) and yields
rQ/3 • = [a, (3; 5],
g-tsTlg =
(B.13.10)
whereas the formulae in eq. (4.16) which define Christoffel’s
symbols of second kind should take the form
+9n+t
’6{nap}-
It follows from eq. (10) that at the points in En+P which belong
to Rn we have
(f^)o = h/M
841 En
B.13 Riemannian subspace Rn in the Euclidean space
and this yields an important relationship

(B.13.11)

Let us proceed to determination of the coefficients qa/3\n+t-


Calculating the scalar product of eq. (9) and r n+s we arrive at
the system of equations
~
9a0\n+t9
„n+t,n+s _______ _ „n+s
~ *oc0 ■ T .
(B.13.12)

Differentiating the relationships


ra ■ rn+s = 0, rp ■ rn+s = 0
with respect to q@ and qa we obtain
ra0 • rn+s = -rQ dqP QTn+s
= ~T0 ■ dq a
'
Using eq. (4.19) we find
Jn+S\r7, / 71 + 8 1-n+tl f» + »\
rQ/3 • r T = rQ l /J7 J U n + tf J 1 pa J

or
ra0 • r”+s = gn+s" [a, ft 7] + gn+s'n+t [a, ft n + t]. (B.
13.13)
The system of equations (12) takes the form
xn+tgn+t’n+s = A7gn+sn,

(B.13.14)
where, for the sake of brevity,
xn+t=qa0\n+t-[oi,0\n + t], A~, = [a,P\ 7]. (B.13.15)

The quadratic form


n+p n+p
53 53 9 PkPm = g PaP0 + 2g ’
km a0 a n+t
paPn+t + gn+t’n+spn+tPn+s k=1 m=1

is the adjoint form of the quadratic form (ds)2, i.e. it is positive


definite. By virtue of Sylvester’s theorem, the determinant
gn+l,n+l gn+l,n+p
an ai2 . • • n
&2 &22 . . • «2 n
1
Unl an2 • •

(B.13.19)

the following row


i)q 5 • • • 5 {9n+t,n)q
{9n+t, (B.13.20)

The ratio of the determinant obtained to determinant (19) is
readily 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
843 En
B.13 Riemannian subspace Rn in the Euclidean space

and to recall that \a\ alu are the cofactors of the elements of
the 7 — 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 (3 of the radius-vector r at the points
a

of Rn
(ra0)o Pap = + {[a, fcn + *]0 - M2+t [a, /?; 7]a) ro+t

VpPa + {[a, fcn + f]0 - Ml+t [a, /?; 7]a) r”+t.

Here
1 &9a,,n+t
dQ/3,n+t ^9a(3 \
[a, f3;n + t]0 = 2 dqh (B.13.22)
dqa dqn+t) Q'

Calculation of the coefficients of the expression of Vj3pa in


terms of the base vectors pa and rg+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 g /3 (a, f3 = 1,... , a

n) are required in the vicinity of Rn belonging to En+P rather than


in Rn, since [a,/3;n +1\ contains the derivatives of gap with
0

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 En+P
b = bara + bn+t rn+t (B.13.23)
the values ba 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
bp = b • r0 = baga0, bn+« = b • r"+« = bn+tqn+t’n+“. (B.13.24)

At the points of En+p belonging to Rn, we have


(bp)0=(ba)oaap, i.e. (ba)Q = aa~< (&7)0 = (6“)0 , (B.13.25)

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 = Y.9n+q’Sh* = gn+q’aba + gn+q’n+tbn+t,
s=1
and the second set of relationships (24) takes the form

9n+q’aba = gn+t'n+q (bn+t - bn+t) ■ (B.13.26)


an «12 9l3
®2 a 22 923
1
031 932 933
B.14 The Riemann-Christoffel tensor 845

we find that
31
n
„31 £o3’
-Ml Wo 9o -Mi =
TT „33>
a #0
Formula (21) takes the form

Pot(3 (B.13.32)

Comparing it with eq. (7.19) we again obtain expressions


(9.16) for the coefficients bgaj3 of the second quadratic form of
the surface.

B.14 The Riemann-Christoffel tensor

Let us consider two consequent covariant differentiations V/?


V7 of the covariant values ca. Recalling that V7ca is a
component of the tensor of second rank which is covariant
with respect to 7 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 expression for V/3V7ca. Thus, we obtain
V0Vyca - = c^R^, (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 7 and /3 and contravariant 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 /?, 7, fi and contravariant with respect to a. This


tensor is referred to as the Riemann-Christoff el tensor. Its covariant
components are expressed in a somewhat simpler form
Qocl'R'yPn- (B.14.3)
It follows from eq. (1) and Ricci’s theorem of Sec.
B.5 that
— V7V^ = (B.14.4)
The expression of the covariant components of the
Riemann-C hr is toffel tensor is transformed with the help of
Christoffel’s symbols of first kind
R'yfii' — 9au 9 [75 &] — 9otu g ^ 9
+ v} gSa [7, g; a] - [7,6; v\ g6a [/?, g; a\.
Then we have
gav-f^ga<7 [7, m; <*] -g^-f^g010 [/?,m;a] = [7, m; H - ^7 [ft w v\ -
OLG r 1 OLG \0 1
9 [7, /x; a] + gaa [/?, /x; a1 dqy ‘
dqP
Expanding the expressions for the derivatives of the brackets
and substituting the following equalities
dgot
v
[a, /?; v\ + [/?, v; a],dga [a, 7; v\ + [7, v\ a],
v
dqP deft
we obtain
= 1 / d2}gyiJ l <92g7/l, d+2g0tl d2gf)v \
2 \d(f d(f dqPdcf depdq1' dq~<dq^ J
gaa ([7, <A [v, ft a\ - [ft m; °\ W, 7;«]) •

From the total number of n4 of the components of the


Riemann-Christoffel tensor there are j^n (n — l)2 independent
2

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) coordinate 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
B.14 The Riemann-Christoffel tensor
847

the Riemann-Christoffel tensor is identically equal to zero,


then the manifold is Euclidean. Alternatively, there can be
assumed such coordinates q1,... , q that the quadratic form
71

(ds)2 , expressed in terms of these, will have constant


coefficients. The non-trivial Riemann-Christ offel tensor tes-
tifies that the manifold under consideration is not Euclidean. If
(ds) 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
V/3V7ca = V7V/3Ca,

(B.14.6)
References

[1] Appell, R 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, Vol.l, 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, R.L., Ashley, H., Halfman, R.L.
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. Ill, 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. Investigations of dynamics of non-
holonomic systems. Gostekhizdat, 1949.
[18] Chaplygin, S.A. On a sphere rolling on a horizontal plane
(in Russian), 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, 379-381.
[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. Legons sur la theorie generate des
surfaces. T.l, paris, 1887.
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.ll, 1947, 301-312.
[29] Gantmakher, F.R., Levin, A.M. Theory of flight of
uncontrolled rockets (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 Russian). 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.I. 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
852 References

[43] Ishlinsky, A.Yu. On the theory of complex systems of


gyroscopic stabilisation (in Russian). Applied
Mathematics and Mechanics, vol. 22, 1958, 359.
[44] Jacobi, C.G.J. Vorlesungen iiber Dynamik. Berlin, 1884.
[45] Jordan, C. Cours d’analyse de l’Ecole Polytechnique, t.3,
Paris, Gauthier-Villars, 1887.
[46] Kantorovich, L.V., Krylov, Approximate methods of
higher analysis (in Russian). Gostekhizdat, 1941.
[47] Kilchevsky, N.A. Elements of tensor calculus and its
application in mechanics (in Russian). Gostekhizdat,
1954.
[48] Kirchhoff, G. Vorlesungen iiber mathematische Physik,
Mechanik. Leipzig, 1876.
[49] Klein, F. Vorlesungen fiber die Entwicklung der
Mathematik im 19. Jahrhundert. Springer, Berlin 1927.
[50] Klein F. und Sommerfeld, A. Uber die Theorie des
Kreisels. Heft 1, Kap. 1, S. 7-68, Leipzig, 1897.
[51] Kochin, N.E. Vector calculus and the basics of tensor
calculus (in Russian). GTTI, 1938.
[52] Krylov, A.N. Sur la variation des elements des or bites
elliptique de planetes, 1915. In: Krylov, A.N Collection of
papers. Publishers of the USSR Academy of Science,
1955, vol.6. 249-266.
[53] Krylov, A.N. Cardan’s suspension in ships (in Russian).
In: Krylov, A.N Collection of papers. Publishers of the
USSR Academy of Science, 1955, vol.12, 193-258.
[54] Krylov, N.M., Bogolybov, N.N. Introduction to nonlinear
mechanics (in Russian). Kiev, 1937.
[55] Lagrange, J.L. Mechanique Analytique. Paris, 1788.
[56] Loizansky, L.G. and Lurie A.I. A course of Theoretical
Mechanics (in Russian). Gostekhizdat, 1955.
[57] Lurie A.I. On the theory of finite rotation of a rigid body
(in Russian). Applied Mathematics and Mechanics, vol.
21, 1957, 371-373.
[58] Lurie A.I. Notes on analytical mechanics (in Russian).
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
Polytechnical 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 Mechanics, 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. 11,
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
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 gyroscope in Cardan’s suspension (in
Russian). Applied Mathematics and Mechanics, vol. 22,
1958, 374-381.
[79] Rumyantsev, V.V. On the stability of motion of a
symmetric gyroscope in Cardan’s suspension , part II (in
Russian). Applied Mathematics and Mechanics, vol. 22,
1958, 499-503.
[80] Saxon, D., Cahn, A.S. Modes of vibration of a
suspended chain. Journal 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). Transactions of the Leningrad
Industrial Institute, 1937, No. 6, 106-112.
[82] Stepanov. A course on differential equations (in
Russian). GITTL, 1945.
[83] St tickler, 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.
References 855

[92] Vyshnegradsky, I.A. 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, I.M. 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.
Index

acceleration precession,
absolute, 89 52, 54 roll, 55
of a particle moving slide, 55 spin, 52,
over the rotating earth, 54 trim, 58
93 addition, 90 angular, velocity roll, 55
81 centripetal, 82 yaw, 55, 58
Coriolis, 90 angles
generalised, 26 of a Euler’s, 51, 54
point in a rigid body, angular
81, 86 momentum
relative, 89, 90 relative, 173
rotational, 82 angular velocity, 130, 138
translational, Darboux’s equation, 142
89, 90 determination of a rigid
of a particle moving over body position, 139, 142
the rotating earth, 92 anomaly
action eccentric, 578
Lagrange’s, mean, 579 true,
737, 776 angle 579 aphelion, 577
ascend, 55 apocentron, 577
attack, 55 apogee, 577 axes
heading, 55 airplane,
heel, 58 54 ship,
nutation, 52, 54 58
pitch, 55
858 Index

velocity, 55 coordinate
cyclic, 371 generalised,
base vector, 805, 23 cyclic, 367 excessive, 28
819 binormal, variation, 34, 36 ignorable,
106 bracket 371 positional, 367, 371
Lagrange, 539, quasi-cyclic, 388
785 transformation, 50
fundamental, coordinates
539 Poisson, curvilinear orthogonal,
537, 786 828 curvature, 106 line, 823
fundamental, normal section of
538 surface, 823 surface line
canonical transformation geodesic, 824
invariance, 550 degrees of
Cardan’s suspension, 62, 67, freedom, 22
125, 129 number, 22
double, 64 derivative
Cayley-Klein’s parameters, covariant, 816 on
135, 136 centre of inertia, surface, 827 derivative of
158 chain hanging with a vector relative (local), 87
mass on the end, 698, 723 determinant, 783
ChristoffePs square brackets, differential of quasi-
177 Christoffel’s symbol first coordinates, 32
kind, 177 second kind, 312, differentiation
813 coefficients covariant, 837
gyroscopic, 310 in Riemannian space,
Lame, 828 837 matrix, 802 of surface
composition of motions, 102 vector, 825 discriminant of
condition of kinematic quadratic form, 791
feasibility of adjacent motion, displacement virtual, 35
692 constraint, 19 force of, double-crank mechanism,
267 holonomic, 19 28 dual vector, 805, 820
non- dyad, 158
stationary, 21
rheonomic, 21 elastic body, 227 element
scleronomic, 21 kinematic, 742 metric,
stationary, 21 742 elements
ideal, 271 non- elliptic orbital, 581
holonomic, 20
one-sided, 20
principle of constraint
release, 267
two-sided, 20
constraint force, 267
generalised, 349, 353,
Index 859

energy of system of particles,


accelerations, 176 278 stable, 280 unstable,
kinetic, 151 amplitude, 718 280 relative, 461
in terms of generalised ve- Euclidean
locities, 151, 152, 171 in subspace, 838
terms of quasi-velocities, Euler’s angles, 51,
155, 171 54
of system subjected to
stationary constraints, 153 finite rotation, 114, 130
potential, 208, 209 commutation, 122
amplitude, 718 generalised, composition, 120 in terms of
211 of attraction force, 217 Euler’s angles, 124
energy integral, 308 Rodrigues-Hamilton’s
equation central parameters, 115
fundamental, 275 subtraction, 122
Lagrange’s, 275 Darboux- foci
Rikatti, 143 Kepler, 579 conjugate kinetic, 677,
of dynamics, 691 force
fundamental, 273 relative aerodynamic resisting,
motion canonical, 459 260 circulatory, 215
equations Coriolis’s, 514 dissipative,
canonical 248 power, 251 elastic, 227
Hamiltonian, 529, 590 generalised
of relative motion, 555 for generalised
variational, 631, 633 coordinate, 203
integration, 634 equations of for quasi-coordinate,
motion 204, 207
AppelPs, 418, 420, 423 in terms of the
Chaplygin’s, 424 Euler- potential energy, 209
Lagrange, 392 gravity, 216 gyroscopic, 310
Lagrange’s generalised, 310 positional,
of the first kind, 270 213, 214 potential, 208, 211
perturbed, 586, 623, indication, 210 radial
637, 649 Routh, 371 correction, 215 resisting,
the Euler-Lagrange, 434 250 thrust, 514 formula
equilibrium Frenet, 107,
in presence of 823 friction
Coulomb’s friction, Coulomb, 249, 254, 296
296 viscous, 249 friction gear,
of a rigid body 429
suspended on elastic rods,
291 of heavy rod, 283
860 Index

frictional pole, 299 perturbed, 625, 642


function kinetic energy
dissipation, 249, in terms of generalised
251 generating, velocities, 171
523, 546 in terms of quasi-
Hamilton’s velocities, 171
characteristic, of relative motion, 175
767, 769 principal, 730, of translational motion, 175
731 Hamiltonian, 528 kinetic potential Routhian,
homogeneous, 154 372 Kronecker symbol, 48
Lagrange, 276
functional Lagrange’s equations of
stationary value, 672 the first art, 270 of the
second kind, 304
generalised momentum, 155 Lagrange’s multipliers,
geodesic line, 823, 825 269 Legendre’s
gravitational field, 216, 288 condition, 686 Levi-
gyroscope Civita symbol, 48
in Cardan’s suspension,
183, 334 Magnus effect, 262
mounted on a moving mathematical
platform, 468 pendulum, 250 double
gyroscopic platform, 188, with a square-law
515 resisting force, 252,
Hamilton’s action, 670, 329
679 minimum, 676 matrix
variation, 671 heavy top addition,
influence of the rotating 787
Earth, 484 adjoint,
helical spring, 245 794 block,
Hooke’s law, 228 784
multiplication,
ideal constraint 792 conform, 788
constraint force, 356 deficiency, 784
indication of potential force, determinant, 783
210 inertia ellipsoid, 168 diagonal
central, 168 inertia tensor, elements, 786
161, 163 transformation, 162 differentiation,
Jacobi’s condition, 686 802 Hess’s, 783
Jacobi’s identity, 538 influence, 231
Jacobian, 784 inverse, 795
Jacobi’s, 782
Keplerian motion, 377 multiplication, 788
Earth satellite by scalar, 787
null, 782 of
inertia, 163 order,
Index 861

rectangular, elliptic, 582


781 rotation, in a resistive medium,
50 square, 598 in homogeneous
782 gravitational field, 770,
skew-symmetric, 778 on a conical surface,
786 symmetric, 757 on developable
786 subtraction, surface, 325 on surface,
787 transpose, 318 on surface of
786 unit, 783 revolution, 324 perturbed,
mean motion, 637 planar, 428 relative,
579 mechanism, 88 rigid body
360 crankshaft, carrying rotating
363 method flywheels, 486
Galerkin’s, 723 free, 438, 440, 442
ignoration of cyclic having a cavity with fluid,
coordinates, 371 492
Ritz’s, 719 relative, 475 rocket, 512
variation of parameters, spinning shell, 260, 443
585 metric stability, 747
element of action, 742 unperturbed stationary, 660
moment of inertia, 161 about kinematically stable, 660
an axis, 161 principal central,
168 momentum, 172 motion nodal axis, 52
Earth satellite, 625, 642 parallel translation of vector,
elastic solid, 498 gyroscopic 826 parameters of Rodrigues-
platform, 515, 519 heavy top, Hamilton, 115
533 carrying a flywheel, 536 path
influence of misbalance, 605 true, 670, 742
in field varied, 670, 682,
of a central force, 742 pendulum
377, 532, 552, 567, mathematical,
568, 576, 756, 772 727 physical, 356
of attraction of two
centres, 760 pericentron, 577
in the gravity field of the perigee, 577
rotating Earth, 590 Keplerian, perihelion, 577
377, 532, 552, 567, 568, 576, perturbation of trajectories,
756, 772 perturbed, 621 747 plane
material system perturbed, tangent contact,
649 natural trihedron, 105 105 Poisson bracket,
particle 537 potential
kinetic,
276 power,
247
862 Index

dissipative forces, 251 rod


in terms of generalised elastic rotating, 707, 726
velocities, 248 virtual, 248 inextensible, 506 heavy, 283
principal angular in an elliptic cup, 286
momentum, 172 principal plain axis, 242 under
axes of inertia, 165 central, bending, torsion and
168 compression, 238 rod
principal normal to a curve, system
106 principle statically determinate,
Hamilton-Ostrogradsky, 231 statically indeterminate,
672, 698 234 Rodrigues formula, 112
of constraint release, rolling, 402
267 of virtual work, 280 without slipping,
stationary action, of Jacobi, 93 ring, 79 sphere,
740 Lagrange, 737 product 79 rotation
dyadic, 158, 789 scalar, about a fixed point, 61
834 earth satellite about its cen-
product of inertia, 161 tre of inertia, 611, 620
flexible shaft, 462 rigid body,
quadratic form, 797 adjoint, 147
796 of surface first, 821 about a fixed point,
second, 821 positive 170, 180
definite, 797 near vertical, 761,
semi-definite, 797 765 with a fixed point,
quasi-coordinates, 483 Routhian function,
32 quasi-velocities, 371, 373
30
safety parabola,
regular precession, 147, 756, 779 shape of
665 resultant force, 205 the earth, 221 shell
rigid body, 19 carrying flywheels
carrying an unbalanced kinetic energy, 184
flywheel, 185 solid, 498 space
in a central force field, Euclidean, 833
288 suspended on elastic Riemannian, 836
rods, 291 with a fixed point, sphere
170, 213 rigidity rolling on a rough
bending, 240 torsional, surface, 193, 395, 425
240 spinning shell, 443 stability
ring of regular precession of
rolling on a rough gyroscope, 665
surface, 425 stability of motion, 747
submatrix, 783
Index 863

Sylvester’s criterion, generalised, 525


153 symbol Euler-Chasles,
Christofell’s first 117 Jacobi’s, 562
kind, 813 second kind, Lagrange-
813 Levi-Civita, 807 Dirichlet, 280
symbols Liouville’s, 549
three-index, 39, 40 Meusnier, 823
system Poisson, 541
four-rod, 343 free, reciprocal Maxwell’s,
19 holonomic, 20 linear 231 Ricci, 817 Stackel’s,
equations solution, 796 571 Steiner’s, 163
Liouville, 574 particle, Sylvester’s, 798 theory
19 prestressed, 292 Sommerfeld’s, 216
three-rod, 281 two three-index symbols, 76
heavy tops, 383 two- trajectory of a particle in the
rod, 339, 364 system of potential field, 316
axes geocentric, 91 transformation
rotation, 127 inertial, 47 canonical, 543, 559
rotation matrix, Legendre’s, 523, 524
50 moving, 47 trihedron
rotation matrix, 50 natural, 106 motion,
109 twist of curve, 107 two-
tangent vector, axle trolley, 42, 196, 400,
105 tensor, 802, 425
834
dyadic representation, variables
810, 834 fundamental, canonical,
810 in non-orthogonal 529 variation
coordinates, 810 asynchronous,
metric, 812, 820 734 coordinate
of second rank synchronous, 36
symmetric, 160 generalised coordinate, 34
Ricci, 654 quasi-coordinate, 39
Riemann- synchronous, 734 vector
Christoffel, 846 skew- angular
symmetric, 811 acceleration, 81
symmetric, 810 theorem angular velocity,
addition of accelerations, 71, 74
90 addition of velocities, components
89 Castigliano, 231 contravariant, 808
covariant, 808 Darboux,
107 geodesic curvature,
823 in Riemannian space,
837 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
moving rigid shell, 490
physical pendulum, 356
rotating rod, 506, 707, 726
work
complementary, 526
elementary, 203, 204, 208 of
potential forces, 209

Вам также может понравиться