You are on page 1of 458

MECHANICS,

TENSORS
&
VIRTUAL WORKS

MECHANICS,
TENSORS
&
VIRTUAL WORKS

Yves R Talpaert
Faculties of Science and Engineering at
Algiers University, Algeria
Brussels University, Belgium
Bujumbura University, Burundi
Libreville University, Gabon
Lom University, Togo
Lubumbashi University, Zaire and
Ouagadougou University, Burkina Faso

CAMBRIDGE INTERNATIONAL SCIENCE PUBLISHING

Published by
Cambridge International Science Publishing
7 Meadow Walk, Great Abington, Cambridge CB1 6AZ, UK
http://www.demon.co.uk/cambsci
First published 2003
Yves Talpaert
Cambridge International Science Publishing

Conditions of sale
All rights reserved. No part of this publication may be reproduced or transmitted in any form or by any means,
electronic or mechanical, including photocopy, recording, or any information storage and retrieval system, without permission in writing from the copyright holder.
British Library Cataloguing in Publication Data
A catalogue record for this book is available from the British Library

ISBN 1 898326 11 8

Production Irina Stupak


Printed by Antony Rowe Ltd, Chippenham, England

About the author

Yves R Talpaert received his MS (1968), Agrgation (1968), and PhD


(1974) degrees in applied mathematics from Brussels University, where he
taught mathematics for a few years. A past Professor at Algiers University,
Algeria and other universities and engineering institutes in Africa, he is a
prolific French author of books on mechanics and differential geometry. He
has written papers on dynamics applied to astronomy where he expounded
an original fluid-dynamical approach, statistical mechanics models, a
variational principle and so on.

PREFACE

This book is intended for third year students in mathematics, physics and engineering
within the context of a first one-semester course in mechanics. Most of the text comes from
this level courses I taught at several universities and engineering schools.
The various chapters connect the notions of mechanics of first and second years with
the ones which are developed in more specialized subjects as quantum physics, continuum
mechanics, fluid-dynamics, special relativity, general relativity, cosmology, meteorology,
electromagnetism, stellar dynamics, celestial mechanics, applied differential geometry and so
on.
This course of Analytical Mechanics synthesizes the basic notions of first level
mechanics and leads to above-mentioned disciplines by introducing various mathematical
concepts as tensor and virtual work methods. A measured and logical progression towards
notions of mathematics and mechanics give this book its originality.
First, the notion of dynam is introduced in Chapter 0, in particular the one of velocity
dynam.
In Chapter 1, the study of Statics is divided into two parts. After recalling the classic
method, the one of virtual work is developed with numerous exercises of classical mechanics.
Tensor theory is very expanded and illustrated in Chapter 2. Tensors are intrinsic
mathematical beings and are suitable for the expression of laws of mechanics (and physics)
regardless of the choice of coordinate system. This property by oneself justifies the extent of
this study of tensors. The reader will better view the notion of inertia tensor within this widen
context.
Devoted to this particular tensor, Chapter 3 especially prepares for the study of
continuum mechanics. Inertia ellipsoid and principal axes are examples among others.
Chapter 4 shows kinetics and dynamics of systems with the help of dynams. But
tensor calculus is also very helpful to write theorems deduced from postulates.
Lagrangian dynamics and variational principles are at the root of analytical mechanics
introduced in Chapter 5. Lagrangian dynamics shows another formulation of motion
equations from the notion of virtual displacements and the profitable use of scalar functions
leads to Lagranges equations. Eulers equation and Hamiltons principle are considered in
the context of variational calculus. Euler-Noether theorem concludes this essential chapter.
Hamiltonian mechanics, which constitutes the last chapter, is dealt with canonical
equations, Lagrange and Poisson brackets, canonical transformations and Hamilton-Jacobi
equation finally. This chapter especially prepares for the formalism of quantum mechanics
and for celestial mechanics notably.

vi

Preface

Many books relating to the developments of tensor theory are either too abstract since
aimed at algebraists only, or too quickly applied to physicists and engineers. I have striven for
bringing closer these points of view; so the various chapters are intended for mathematicians
(which will find an illustrated presentation of mathematical concepts and solved problems)
and for physics and engineering students too, since the mathematical foundations are basically
introduced in a practical way.
As well as tensors, the virtual work concept is systematically used. Being clearly
introduced from virtual displacements (and virtual velocities), this notion plays an essential
role in continuum mechanics.
The two previous mathematical tools give the considered mechanics subjects a great
unity and must be known by every mathematician, physicist and engineer. And after all, they
make the reading of my previous book treating of Differential Geometry easier, which is even
better!
The present book lets overcome the following difficulty. There is often a gap between
academic cycles. Two essential reasons among others are responsible for that. The one
follows from the diversity of teaching establishments of first and second years. The second is
due to the different nature of academic cycles, namely: general courses at first and second
years, specialized courses later.
One of my goals has been to reduce this gap, the intensive use of tensor calculus
contributes to that; this realization facilitates the writing of a second volume which will deal
with rigid bodies, perturbations and continua.
All the proofs and 78 solved exercises are detailed. The important propositions and the
formulae to be framed are shown by  and .
In writing this new book, I had the following assertion fresh in my mind: Pedagogy
contributes to Rigor.

Acknowledgements. I am grateful to Professor Michel N. Boyom (Montpellier University) for


a critical reading of chapters.
Many thanks to my former students who let me expound on the material that resulted in this
book.

Yves Talpaert

CONTENTS
PREFACE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
CONTENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii
Chapter 0.

Chapter 1.

REQUIREMENTS

1.

POINT SPACE AND VECTOR SPACE . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.1

Point space (or affine space) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2

Frame of reference and basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

2.

DYNAMS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

2.1

Dynam definition and reduction elements . . . . . . . . . . . . . . . . . . . . . . . . . 4


Dynam definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
Representation of dynams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2.2

Properties and operations on dynams . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6


Equality of dynams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
Operations on dynams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
Equiprojective fields of moments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
Invariants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
Reduction of a vector system and dynam . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.3

Dynam of velocities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
Velocity field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
Dynam of velocities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

2.4

Acceleration vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

2.5

Sliding velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

3.

EXERCISES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .27

STATICS

33

1.

CLASSIC METHOD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

1.1

Mechanical actions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
Moment of a force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
Dynam of a mechanical action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

1.2

Classification of forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
External forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
Internal forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

1.3

Equilibrium conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
Definitions and conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
Particular collections of forces applied to a rigid body . . . . . . . . . . . . . . . . . 43

vii

viii

Contents
1.4

Types of equilibrium of rigid bodies and structures . . . . . . . . . . . . . . . . 45

1.5

Stress and contact dynam . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .


Stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Contact dynam . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Dry friction and Coulomb laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.6

Types of constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
Punctual constraint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
Rectilinear constraint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
Annular-linear constraint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
Ball-and-socket joint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
Plane support . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
Sliding pivot . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
Sliding guide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
Screw joint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
Pivot . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
Embedding or welded joint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

1.7

Free-body diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

2.

METHOD OF VIRTUAL WORK . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .70

2.1

Number of degrees of freedom and generalized coordinates . . . . . . . . .


Number of degrees of freedom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Generalized coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Types of constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.2

Virtual displacements and virtual velocities . . . . . . . . . . . . . . . . . . . . . . . 78


Generalized coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
Definition and expression of virtual displacements . . . . . . . . . . . . . . . . . . . 79
Virtual velocity and examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
Virtual fields and dynams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . 91

2.3

Virtual work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
Definitions, rigid body and ideal constraint . . . . . . . . . . . . . . . . . . . . . . . . . 94
Principle of virtual work (First expression) . . . . . . . . . . . . . . . . . . . . . . . . . 98
Principle of virtual work (Second expression) . . . . . . . . . . . . . . . . . . . . . . 106

3.

EXERCISES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

Chapter 2.

TENSORS

47
47
48
50

71
71
72
75

135

1.

FIRST STEPS WITH TENSORS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

1.1

Multilinear forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135


Linear mapping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
Multilinear form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

1.2

Dual space, vectors and covectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137


Dual space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
Expression of a covector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
Einstein summation convention . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
Change of basis and cobasis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . 140

1.3

Tensors and tensor product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

Contents

ix

Tensor product of multilinear forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143


Tensor of type (10 ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

(10 )
Tensor of type ( 02 )
Tensor of type ( 02 )
Tensor of type (11 )
Tensor of type

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150

Tensor of type (qp ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151


Symmetric and antisymmetric tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
2.

OPERATIONS ON TENSORS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156

2.1

Tensor algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156


Addition of tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
Multiplication by a scalar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
Tensor multiplication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

2.2

Contraction and tensor criteria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158


Contraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
Tensor criteria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162

3.

EUCLIDEAN VECTOR SPACE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

3.1

Pre-Euclidean vector space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164


Scalar multiplication and pre-Euclidean space . . . . . . . . . . . . . . . . . . . . . . 164
Fundamental tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

3.2

Canonical isomorphism and conjugate tensor . . . . . . . . . . . . . . . . . . . . 166


Canonical isomorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
Conjugate tensor and reciprocal basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
Covariant and contravariant representations of vectors . . . . . . . . . . . . . . . 170
Representation of tensors of order 2 and contracted products . . . . . . . . . . 172

3.3

Euclidean vector spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174

4.

EXTERIOR ALGEBRA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178

4.1

p-forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
Definition of a p-form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
Exterior product of 1-forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
Expression of a p-form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
Exterior product of p-forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
Exterior algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185

4.2

q-vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188

5.

POINT SPACES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192

5.1

Point space and natural frame . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192


Point space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192
Coordinate system and frame of reference . . . . . . . . . . . . . . . . . . . . . . . . . 192
Natural frame . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194

5.2

Tensor fields and metric element . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197


Transformations of curvilinear coordinates . . . . . . . . . . . . . . . . . . . . . . . . 197

Contents
Tensor fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
Metric element . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
5.3

Christoffel symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203


Definition of Christoffel symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203
Ricci identities and Christoffel formulae . . . . . . . . . . . . . . . . . . . . . . . . . 206

5.4

Absolute differential, Covariant derivative, Geodesic . . . . . . . . . . . . .


Absolute differential of a vector and covariant derivatives . . . . . . . . . . . .
Absolute differential of a tensor and covariant derivatives . . . . . . . . . . . .
Geodesic and Eulers equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Parallel transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Absolute derivative of a vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.5

Volume form and adjoint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216


Volume form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
Adjoint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218

5.6

Differential operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Gradient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Divergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Curl . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Laplacian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

230

EXERCISES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

232

6.

Chapter 3

MASS GEOMETRY AND INERTIA TENSOR

207
207
209
211
213
214

220
220
225
228

263

1.

MASS DISTRIBUTION AND INTEGRALS . . . . . . . . . . . . . . . . . . . . 263

1.1

Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

263

1.2

Integrals of real-valued functions and vector functions . . . . . . . . . . .

265

2.

CENTER OF MASS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267

2.1

Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.2

Subdivision . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269

2.3

Theorems of Guldin (and Pappus) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271

3.

INERTIA TENSOR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272

3.1

Moments and products of inertia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272

3.2

Inertia tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274

4.

INERTIA ELLIPSOID . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 276

4.1

Moment of inertia about an axis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4.2

Equation of the quadric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 278

4.3

Nature of the quadric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279

4.4

Radius of gyration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280

5.

PRINCIPAL AXES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281

267

276

Contents

Chapter 4

Chapter 5

xi

5.1

Fundamental theorem about a symmetric tensor . . . . . . . . . . . . . . . . . 281

5.2

Equal eigenvalues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284

5.3

Inertia ellipsoid and principal axes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 286

5.4

Material symmetries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287

6.

THEOREM OF STEINER (and HUYGENS) . . . . . . . . . . . . . . . . . . . . 288

7.

EXERCISES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 291

KINETICS AND DYNAMICS OF SYSTEMS

299

1.

NEWTONS POSTULATES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299

1.1

Experimental laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 300

1.2

Postulates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301

1.3

Galilean relativity and inertial frames . . . . . . . . . . . . . . . . . . . . . . . . . . 303

2.

KINETICS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307

2.1

Kinetic dynam . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307

2.2

Kinetic energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 308

3.

THEOREMS OF MECHANICS OF SYSTEMS . . . . . . . . . . . . . . . . . . 309

3.1

First integrals of a system of particles . . . . . . . . . . . . . . . . . . . . . . . . . . 309

3.2

Linear momentum theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 310


Linear momentum theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 310
Theorem of conservation of mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311
Theorem of motion of mass center . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311
Special case of rigid bodies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312

3.3

Angular momentum theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 316


Angular momentum theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 316
Relation between kinetic dynam and dynam of forces . . . . . . . . . . . . . . . . 317
Conservation of angular momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 318
Special case of rigid bodies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319

3.4

Kinetic energy theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323


Kinetic energy theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323
Special case of rigid bodies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 326

4.

EXERCISES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331

LAGRANGIAN DYNAMICS AND VARIATIONAL PRINCIPLES 339


1.

LAGRANGIAN DYNAMICS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 339

1.1

Holonomic and scleronomic systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 340

1.2

DAlembert-Lagrange principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342

xii

Chapter 6

Contents

1.3

Lagranges equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Lagranges equations in the general case . . . . . . . . . . . . . . . . . . . . . . . . .
Lagranges equations for conservative forces . . . . . . . . . . . . . . . . . . . . . .
Lagranges equations with undetermined multipliers . . . . . . . . . . . . . . . .

1.4

Configuration space and Lagranges equations . . . . . . . . . . . . . . . . . . 354

1.5

Adjoint Lagrangian and first integrals . . . . . . . . . . . . . . . . . . . . . . . . . 358

2.

VARIATIONAL CALCULUS AND PRINCIPLES . . . . . . . . . . . . . . . 360

2.1

Eulers equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361


A variational problem and variations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361
Eulers equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 364

2.2

Hamiltons variational principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 368


Hamiltons postulate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 368
Hamiltons principle and motion equations . . . . . . . . . . . . . . . . . . . . . . . . 369

2.3

Jacobis form of the principle of least action of Maupertuis . . . . . . . . 371

3.

EULER-NOETHER THEOREM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 374

3.1

One-parameter group of diffeomorphisms . . . . . . . . . . . . . . . . . . . . . . 374

3.2

Euler-Noether theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 376

3.

EXERCISES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 379

HAMILTONIAN MECHANICS

344
344
348
350

393

1.

N-BODY PROBLEM AND CANONICAL EQUATIONS . . . . . . . . . 393

2.

CANONICAL EQUATIONS AND HAMILTONIAN . . . . . . . . . . . . . 397

2.1

Legendre transformation and Hamiltonian . . . . . . . . . . . . . . . . . . . . . 397

2.2

Canonical equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 401

2.3

First integrals and cyclic coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . . 404

2.4

Liouvilles theorem in statistical mechanics . . . . . . . . . . . . . . . . . . . . . . 406

3.

CANONICAL TRANSFORMATIONS . . . . . . . . . . . . . . . . . . . . . . . . . 409

3.1

Lagrange and Poisson brackets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 409


Preservation of canonical form and Poisson bracket . . . . . . . . . . . . . . . . . . 409
Poisson bracket and symplectic matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . 411
Lagrange and Poisson brackets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 415

3.2

Canonical transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 416


Canonical transformations and brackets . . . . . . . . . . . . . . . . . . . . . . . . . . . 417
Canonical transformations and generating functions . . . . . . . . . . . . . . . . . 419

4.

HAMILTON-JACOBI EQUATION . . . . . . . . . . . . . . . . . . . . . . . . . . . 425

4.1

Hamilton-Jacobi equation and Jacobi theorem . . . . . . . . . . . . . . . . . . 425

4.2

Separability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 430

5.

EXERCISES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 439

Contents

xiii

BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 453
INDEX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 455

CHAPTER 0

REQUIREMENTS

First, let us recall the mathematical notion of point space (also called affine space)
from the one of vector space.

1. POINT SPACE AND VECTOR SPACE


1.1

POINT SPACE (OR AFFINE SPACE)

In mechanics, besides vector spaces it is necessary to consider spaces consisting of


points. They can be customary as the classic 3-dimensional Euclidean space or very helpful to
mechanics as phase spaces.
Let E be a (real) vector space.
With this vector space, we can associate a point space denoted by E.
How?
There exists a mapping

:EE E
such that:
(i)

r E , a E , !b E : (a, b) = r ,

(ii)

a, b, c E : (a, b) + (b, c) = (a, c) .

Any vector (a, b) is denoted by ab.


Now, given an arbitrary point o E , we define the following mapping:

o : E E : x  o ( x) = (o, x) = ox .
Property (i) implies that, given an arbitrary point o E , the mapping o is a bijection that
establishes the connection between the vector space E and the point space E.
1

Chapter 0

So, we can clearly define the notion of point space:


D

A set E is said to be a point space or an affine space if there is a mapping which


connects pairs of elements of E with vectors of E; namely:

: E E E : (a, b)  ab
such that:
(i) ab = ba ,
(ii) ab + bc = ac ,
(iii) Given arbitrary o E , with any r E is associated only one x E
such that ox = r .
Each element of E is called a point.
D

The dimension of a point space is the dimension of the corresponding vector space.

1.1

FRAME OF REFERENCE AND BASIS


We assume E of finite dimension.

Let B = (e1 ,..., e n ) be a basis of E,


o be a point of E.
The bijection o1 permits to associate a frame of reference of E with the basis B of E. How?
We consider the different points of E defined as follows:

ai = o1 (e i )

i = 1,..., n

or similarly from:

o ( ai ) = e i .

Fig. 1
D

A frame of reference of point space E is the set of n + 1 (affinely) independent points


o, a1 ,..., a n .

So, a frame of reference of a point space E is defined from a reference point o and a basis (e i )
of the associated vector space E.

Requirements

It is denoted by
{ o; oa1 ,..., oa n }

or simply in general:

R { o; e ,..., e }.
1

Now, let us recall the notion of position vector of any point.


Let r be a vector of E,
x be the point of E such that x = o1 (r ) .
The vector r = ox is expressed with respect to B as
n

r = x i e i = ox
i =1

where the various x i are called the components of r relative to B or the coordinates of x
with respect to R. It is the well-known position vector of x with respect to R.
To conclude this section, we recall:
D

A vector space is said to be pre-Euclidean if it is provided with a scalar multiplication.


It is said to be Euclidean if the scalar multiplication is positive definite; that is, given
any nonzero x E , the product scalar x . x = x

is strictly positive.

A point space is said to be pre-Euclidean (resp. Euclidean) if the associated vector


space is pre-Euclidean (resp. Euclidean).

2. DYNAMS
The theory of dynams has been notably developed by German researchers in
mechanics and the term dynam corresponds to the so-called torseur in French. However,
the reader will be careful: the dynam terminology is not unique and has not a universal
acceptation.
The dynams let unify the presentation of the reduction of systems of vectors (as systems of
forces in statics), simplify the expression of theorems of dynamics, express the notion of
velocity fields in kinematics of rigid bodies and so on.
In mechanics, it is interesting to create mathematical beings defined from the resultant and the
total moment (or total torque) of any system of (sliding or bound) vectors by forgetting this
system of vectors and by referring to only previous resultant and moment. These beings are
called dynams.1
First, we are going to give the definition and properties of dynams and finally we will
introduce the notion of velocity dynam.
1

In this section, systems of vectors will be assumed made up of sliding or bound vectors, by knowing that forces
are considered as bound vectors, but as sliding vectors only in the context of unstretchable material systems.

2.1

Chapter 0
DYNAM DEFINITION AND REDUCTION ELEMENTS

2.1.1 Dynam Definition

The equivalence of systems of vectors is a well-known notion developed in first


courses of mechanics; this is essentially recalled for notation.
Let us denote by R(F) the resultant of a system F of (sliding or bound) vectors and by
M a ( F ) the moment of F about a point a.
The equivalence of two systems of vectors is expressed as follows:
R( F ) = R(G ) ,

F ~G

a E : M a ( F ) = M a (G ) .

In this definition it is not necessary to state the second condition at every point a E . It is
sufficient to specify this condition at only one point, since if this condition is fulfilled at b E
for instance, then it is verified at every point a because
M a ( F ) = M b ( F ) + ab R( F ) .
We have thus:
R( F ) = R(G ) ,

F ~G

b E : M b ( F ) = M b (G ) .

So, we say:
D

Two systems of (sliding or bound) vectors are equivalent if they have the same
resultant and the same total moment at an arbitrary point.

The equivalence relation which follows from this definition allows the partition of systems of
vectors into equivalence classes. Indeed, this definition leads to an equivalence relation on the
set C of systems of vectors; that is:
F , G , H C :

F~F ,
F~G

G~F ,

[ F ~ G and G ~ H ]

F~H .

The equivalence class containing the system of vectors F is defined by


D = {S C

S ~ F }.

Now, we can set:


D  A dynam is an equivalence class of systems of vectors.
So, every system of vectors of the equivalence class represents the dynam.
The quotient C / ~ is the set of equivalence classes (dynams).

Requirements

Since all systems of vectors in an equivalence class are characterized by the same resultant
and the same (total) moment at an arbitrary point of E, the definition of dynam can be
expressed as follows:
D  A dynam D is defined by a set of two vector fields:
The one is
E E : a  R( D) ,

which is a mapping such that the image of any point a is the same free vector R.
The other is
E E E : a  (a, M ( D)) = M a ( D) ,
which is a mapping such that the image of any point a fulfills the condition:
a, b E : M a ( D) = M b ( D) + ab R( D) .
D

(0-1)

The vector R(D) is called the resultant of the dynam D (or vector of D) and M a (D)
is the moment of the dynam D at a.
These characteristics of the dynam D are said to be the elements of reduction of D
at the point a. There are also called the components of the dynam.

The dynam is denoted by specifying its elements of reduction as follows:


R
D= .
M a

Remarks.
(i)

The elements of reduction are obviously of different nature.

(ii)

If the elements of reduction of a dynam are determined at a point, then there are know at
every point. The vector field defined by the resultant R( D) is uniform while the vector
field of moments is, of course, not of this type.

(iii) There is a bijection between the set of dynams and the set of pairs made up of the
resultant and moment. Later we will introduce operations on the set of dynams so that
this bijection will be an isomorphism between vector spaces.
Example. Given an orthonormal basis (i , j , k ) of a vector space E, let us consider the vector
field
E E : a  w a = (2 + y + l 2 z ) i + (l z x) j (4 x + l y ) k

where l is a real parameter and a the point of coordinates ( x, y, z ) .


Determine l in order that this vector field is the field of moments of a dynam and find the
resultant of the corresponding dynam.
Answer. Let b( x + X , y + Y , z + Z ) be any point of E.
From
w b = [2 + ( y + Y ) + l 2 ( z + Z )] i + [l ( z + Z ) ( x + X )] j [4( x + X ) + l ( y + Y )] k ,

Chapter 0

we deduce:
w b w a = (Y + l 2 Z ) i + (l Z X ) j (4 X + lY ) k .
The problem is to find a (uniform) vector field
R = ri + s j + uk

such that
w b w a = ba R ,
that is:
i

(Y + l Z ) i + (l Z X ) j (4 X + lY ) k = X Y Z
r s u
2

Y + l 2 Z = uY + sZ ,

l Z X = uX rZ ,
4 X lY = sX + rY

[ u = 1, l 2 = s = 4 , r = l ].

In conclusion, a dynam D is defined for l = 2 and l = 2 ; its resultant is


for l = 2 :
for l = 2 :

R(D) = 2 i + 4 j k ,
R(D) = 2 i + 4 j k .

2.1.2 Representation of Dynams

We are going to introduce the representation of any dynam from the following:
PR1

Every system of vectors defines a dynam, the converse is false.

Proof. First, given a system of vectors, say F, there exists the resultant R(F ) and the
moment M a (F ) of the system F at an arbitrary point a. These elements define a dynam D
such that:
R( F ) = R( D) ,
a E : M a ( F ) = M a ( D) .
Conversely, a dynam does not (completely) define a system of vectors. This is obvious
because equivalent systems of vectors, but which are different, generate the same dynam.
D

The system of vectors F is said to be a representation of the dynam D.

2.2

PROPERTIES AND OPERATIONS ON DYNAMS

2.2.1

Equality of Dynams

Two dynams D1 and D2 are equal if their elements of reduction are equal at every
point.

Requirements

The definition is denoted by

R( D1 ) = R( D2 ) ,
D1 = D2

a E : M a ( D1 ) = M a ( D2 ) .

By using (0-1), we have immediately:


R( D1 ) = R( D2 )
D1 = D2

p E : M p ( D1 ) = M p ( D2 ) .

Remark. Two dynams are equal iff the moments of each dynam about every point are equal:

D1 = D2

a E : M a ( D1 ) = M a ( D2 ) .

The necessary condition is obvious.


The sufficient condition:
a E : M a ( D1 ) = M a ( D2 )

R( D1 ) = R( D2 )

is easily proved. Indeed, a, b E , the equality of the two following equations


M a ( D1 ) = M b ( D1 ) + ab R( D1 ) ,
M a ( D2 ) = M b ( D2 ) + ab R( D2 )
and the equality
M b ( D1 ) = M b ( D2 )
imply:

ab [ R( D1 ) R( D2 )] = 0
that is:
R( D1 ) = R( D2 ) .
D

The dynam zero, denoted D = 0 , is defined by


R( D) = M a ( D) = 0 .

We note that every system of vectors which is equivalent to zero is a representation of dynam
zero.
2.2.2

Operations on Dynams

2.2.2a Addition
D

R
R
The dynam sum of dynams D1 = 1 and D2 = 2 is the dynam
M1 a
M 2 a
R + R2
D1 + D2 = 1
.
M1 + M 2 a

Chapter 0

This is well defined because we say:


PR2

If the systems of vectors { w1 ,..., w n } and { w n +1 ,..., w n+ s } represent the dynams D1


and D2 respectively, then the system of vectors { w1 ,..., w n+ s } represents the dynam
D1 + D2 .

Proof. We have:
n

R( D1 ) = w i ,

R ( D2 ) =

n+ s

wi

i = n +1

i =1

and thus
n+ s

R( D1 ) + R( D2 ) = w i .
i =1

In the same manner, given an arbitrary q E and any ai E belonging to the line of action of
each w i , we have:
n

M q ( D1 ) = qa i w i ,
i =1

M q ( D2 ) =

n+ s

qa i wi

i = n +1

and thus
n+s

M q ( D1 ) + M q ( D2 ) = qa i w i .
i =1

Let us denote the set of dynams by D.


Therefore, we define:
D

The addition of dynams is the inner law


D D D : ( D1 , D2 )  D1 + D2
where D1 + D2 is the dynam sum.

2.2.2b Multiplication by a Scalar


D

R
The product of a dynam D = by a scalar k is the dynam denoted kD such that:
M a
kR
kD = .
kM a

This is well defined because we say:


PR3

If the system of vectors { w1 ,..., w n } represents the dynam D, then the system of
vectors { kw1 ,..., kw n } represents the dynam k D .

Proof. We have:

Requirements

R( D) = w i
i =1

and thus
n

k R( D) = kw i .
i =1

In the same manner, given an arbitrary q E and any ai E belonging to the line of action of
each w i , we have:
n

M q ( D) = qa i w i
i =1

and thus
n

k M q ( D) = qa i k w i .
i =1

Therefore, we define:
D

The multiplication of a dynam by a scalar is the (external) law


R D : (k , D)  k D

where k D is the product of the dynam D by the scalar k.


Of course, we note that the definitions do not depend on the choice of q E .
Remark. The set D provided with the two previous operations is a 6-dimensional vector
space.

2.2.2c Multiplication of Dynams


D

The product of two dynams, denoted by D1 .D2 , is the real

D1 . D2 = R( D1 ) . M a ( D2 ) + R( D2 ) . M a ( D1 ) ,
where a is an arbitrary point.
We note this definition is actually independent of the choice of a E . Indeed, a, b E :
R( D1 ) . M a ( D2 ) + R( D2 ) . M a ( D1 )
= R( D1 ) . M b ( D2 ) + R( D1 ) . ab R( D2 ) + R( D2 ) . M b ( D1 ) + R( D2 ) . ab R( D1 )
= R( D1 ) . M b ( D2 ) + R( D2 ) . M b ( D1 ) ,

this last equality following from the following


R( D1 ) . ab R( D2 ) = R( D2 ) . ab R( D1 ) .
Therefore, we define:

10

Chapter 0

The multiplication of dynams is the law


D D R : ( D1 , D2 )  D1 . D2
where D1 .D2 is the product of dynams D1 and D2 .

Now, let us establish the expression of the product of two dynams in an orthonormal basis of
a vector space E.
From
R( D1 ) = X 1i + Y1 j + Z1k ,

R( D2 ) = X 2 i + Y2 j + Z 2 k ,

M a ( D1 ) = L1i + M 1 j + N1k ,

M a ( D 2 ) = L2 i + M 2 j + N 2 k ,

we deduce the expression of the product of dynams:


D1 . D2 = X 1 L2 + Y1 M 2 + Z1 N 2 + X 2 L1 + Y2 M 1 + Z 2 N1 .
Remark. The reader will verify that the properties of the product of dynams are those of the
inner product of vectors except for one of them, namely:

PR4

The square of a nonzero dynam is not necessarily positive.

Proof. The square


( D1 ) 2 = D1 . D1 = 2( X 1 L1 + Y1 M 1 + Z1 N1 )

can possibly be negative.


We conclude this section by the following
D

Any two dynams are said to be orthogonal if their product is zero.

2.2.3 Equiprojective Fields of Moments

PR5

The field of moments of any dynam is equiprojective.

Proof. Given arbitrary points a and b of E, we have immediately:


ab . M a ( D) = ab . M b ( D)
and so, the projections of M a (D) and M b (D) onto ab being equal, the field of moments is
said to be equiprojective.
Conversely, let us prove the following:
PR6

Every equiprojective vector field is the field of moments of a dynam.

Proof. By the hypothesis, there are vectors M a and M b such that a, b E :


ab . M a = ab . M b

11

Requirements
that is a, b E :
ab . ( M a M b ) = 0 .

- The proof of the proposition is obvious if M a = M b . In this case the resultant of the dynam
is zero.
- Let us consider the general case where the equiprojective vector field is not uniform as
previously.
Let m a , m b and m c be vectors of the equiprojective field at three non collinear points. We
have:
m a . ab = m b . ab ,

(m a m b ) . ab = 0

which means that m a m b is orthogonal to ab.


Similarly, we deduce that m c m b is orthogonal to bc and m c m a is orthogonal to ac.
For obtaining a field of moments, we must find a free vector R such that:
m b m a = ba R ,

m c m b = cb R ,

the third equality m c m a = ca R being automatically verified.


Let us note that R must be orthogonal to the plane determined by the vectors m b m a and
m c m b ; we have thus:
R = k (m b m a ) (m c m b )
= k (ba R) (m c m b )
= k [ba .(m c m b )] R [ R .(m c m b )] ba
= k [ba . (m c m b )] R
which implies:
k=

ba . (m c m b )

The vector R is so determined.


Since it is immediately proved that an equiprojective field is determined by three of its
vectors at three non collinear points, the proposition is thus proved, the dynam being, q E :
R
m .
q

2.2.4

Invariants

An invariant of a dynam is a quantity independent of the point where it is calculated.

(i) The resultant R(D) of a dynam D is a vector invariant, it is a uniform vector field.

12

Chapter 0

(ii) The scalar product


R( D) . M q ( D)

is a scalar invariant. Indeed, it is independent of the choice of q:


a, b E :

M a ( D) = M b ( D) + ab R( D)

and thus:
M a ( D) . R( D) = M b . R( D) .
(iii) Another invariant has already been encountered; namely: the product of dynams.

2.2.5

Reduction of a Vector System and Dynam

An essential problem consists in determining the simplest system of the set of


equivalent systems of (bound and sliding) vectors which define a given dynam; it is the wellknown problem of the reduction of a vector system.
Let R(D) and M q (D) be the elements of the dynam D at q E .
We are going to classify the dynams associated with the different simplest systems of vectors
following from the reduction.
The classification will be made from the scalar invariant R( D) . M q ( D) .
2.2.5a First Case: R( D) . M q ( D) = 0 .
This scalar product can vanish in three cases:
(i) R( D) = M q ( D) = 0 .
These conditions define a dynam associated with a vector system equivalent to zero. It is the
zero dynam denoted by
0
0 .
q
(ii) R(D) = 0 , M q (D) 0 .
These conditions define a dynam associated with a couple of vectors.
First, we recall the following:
PR7

A system of vectors, say F, defining a dynam D is equivalent to a couple iff


[ R( D) = 0 and q E : M q ( D) 0 ] .

Proof. The necessary condition is obvious.


Let us prove the sufficient condition, namely:
If [ R( D) = 0 and q E : M q ( D) 0 ], then the system F is equivalent to a couple.

13

Requirements

Indeed, let M be a nonzero free vector. Let us denote a nonzero (bound) vector located at q by
(q, M ) . Let us search the simplest system of vectors defining a dynam D such that:
R(D) = 0 ,

M q ( D ) = ( q, M ) .

We locate, at a E , the (bound) vector (a, v ) of F such that its moment about q is (q, M ) .
Both vectors (a, v ) and (q,v ) form a system of vectors such that:
R( D) = 0 ,

M q ( D ) = ( q, M ) ;

it is really a couple of vectors.


D

The dynam defined by a couple of vectors is called a couple.

It is denoted:
0
C = .
M
It is a dynam such that the vector invariant (the resultant) is zero and the moment is uniform
(free vector).
(iii) R( D) 0 (with R( D ) . M q ( D ) = 0 ).
These conditions define a dynam of which we are going to express the simplest associated
system of vectors.
First, we recall the following
PR8

A system of vectors defining a dynam D is equivalent to a simple nonzero vector iff


[ q E : R( D ) . M q ( D ) = 0 and R( D) 0 ].

Proof. The necessary condition


F ~ {v}

[ q E : R( D ) . M q ( D ) = 0 and R( D) 0 ]

is immediate since
{v}.{v} = 2 R ({v}) . M q ({v}) = 0
= 2 R( D) . M q ( D)

and
R( D) = R ({v}) = 0 .

Let us prove the sufficient condition, namely:


[ q E : R( D) . M q ( D ) and R( D) 0 ]

Let us consider a E such that


qa =

R( D ) M q ( D )
R 2 ( D)

This vector is orthogonal to R(D) and M q (D) .

F ~ {v}.

14

Chapter 0

We have:
qa R( D) =

1
2

[ R 2 ( D) M q ( D) ( R( D) . M q ( D)) R( D)]

R ( D)
= M q ( D)

So, the system made up of the only vector v = (a, R( D)) has really R(D) as resultant and
M q (D) as moment about q.
In conclusion, the simplest system of vectors associated with D is equivalent to one (bound or
sliding) vector such that the line of action is through q E if M q (D) = 0 or through a E if
qa R( D) = M q ( D) .

A sliding dynam is a dynam such that the resultant is different from zero and the
scalar invariant vanishes.1

It is a dynam of which the moment vanishes at least at a point. We denote any sliding dynam
by S.
PR9

There is a straight line of points about which the moments of a sliding dynam vanish.

Proof. Let a E such that M a (S ) = 0 . There exists a point q such that qa is parallel to R(S ) .
We have thus
M q (S ) = 0
and so there is a straight line of points about which the moments vanish; elsewhere, the
moment is orthogonal to the resultant.
D

The straight line of points about which the moments vanish is called the axis of the
sliding dynam.

We sum up the previous results as follows:


PR10 Any nonzero dynam such that the scalar invariant vanishes is either a couple or a
sliding dynam.
It is a sliding dynam if the resultant is different from zero.
2.2.5b General Case: R( D) . M q ( D) 0 .
We know the Poinsot theorem of reduction:
PR11 Every system of (bound or sliding) vectors is equivalent to a single vector through an
arbitrary point q E , plus a couple.
The point q is called the center of reduction.
It is the general representation of a dynam D of which the single vector is the resultant R(D)
and the moment of the couple is the moment of dynam M q (D) .
1

In French, a sliding dynam is called glisseur.

15

Requirements

Remark. The reduced system of vectors is a system of bound or sliding vectors. So, the
moment of the couple is not considered as a free vector in this study; likewise, the resultant is
not viewed as a free vector but as a (bound) vector located at q or a sliding vector through q.
We are going to show that the set of three vectors of the previous reduced system, which
represents a dynam D, can be expressed in a more interesting form by choosing the center of
reduction suitably.
We say:
D

The central axis of a system of vectors is the set of points where the resultant and the
(total) moment of the system are parallel.

First, we note that:


a, b E : M a ( D) = M b ( D)

ab R(D) = 0

and in this case, ab is parallel to R(D) for any points a and b.


From this remark, if we find a point where the (total) moment is parallel to the resultant, then
the central axis is immediately determined.
We recall that the central axis equation is easily obtained.
Let s be a point of the central axis.
The hypothesis
R( D) M s ( D) = 0
is written q E :
R( D) M q ( D) + R( D) ( sq R( D)) = 0

R( D) M q ( D) + R 2 ( D) sq ( R( D). sq ) R( D) = 0 .
We simplify this equation by particularizing the point s so as to cancel the third term; it is
sufficient to choose s in the plane orthogonal to R(D) through q.
Therefore the equation becomes:
qs =

R( D) M q ( D)
R 2 ( D)

Any point a of the central axis is located by the vector:


qa = qs + sa ;

that is, by choosing q at the origin o of the frame of reference:

r=

R( D) M o ( D)
R 2 ( D)

+ k R( D) .

It is the well-known equation of the central axis of a vector system.


Now, we say the Poinsot theorem of reduction as follows:

(0-2)

16

Chapter 0

PR12 Every dynam D such that the scalar invariant is different from zero is the sum of a
couple C and a sliding dynam S.
At every point of the central axis, the moment of the couple and the resultant of the
sliding dynam are parallel.
We note that the moment M q (D) is the sum of M q (D) parallel to R(D) and M q (D)
perpendicular to R(D) . Therefore, the dynam
R
R
0
D = = +
M q M q M q

is the sum of a couple the moment of which is parallel to the resultant and a sliding dynam
(since R . M = 0 ).
The sliding dynam is represented by the only vector R along the central axis (for any point a
of the central axis, we have M a = 0 ).
PR13 The moment of the couple is minimum at every point of the central axis of the system
of vectors.
Proof. Let a be any point of the central axis,
q be an arbitrary point.
The vector invariant R(D) and the scalar invariant such that:
M a R( D) = M a ( D) . R( D)
= M q ( D ) . R( D )

really imply:
M a M q (D) .

In addition, from
M a ( D) =

M q ( D) . R( D)
R( D )

we deduce for every point a of the central axis the following relation:

M a ( D) =

M q ( D) . R( D)
R 2 ( D)

R( D) .

To conclude this paragraph, we sum up the classification in the following manner:


(i) R(D) 0 :
q E , if R( D) . M q ( D) 0 : general (nondegenerate) dynam,
R( D) . M q ( D) = 0 : sliding dynam.

(ii) R(D) = 0 :
q E , if M q (D) 0 : couple
if M q (D) = 0 : zero dynam.

(0-3)

17

Requirements
2.3

DYNAM OF VELOCITIES

We consider a rigid body B at rest with respect to a frame. There is an infinity of such
frames.
D

This set of frames, in which B is at rest, defines the space connected with the rigid
body as opposed to the frame of reference which is considered as fixed most of the
time.

Let R e = { o; e1 , e 2 , e3 } be an orthonormal frame fixed in E,


R E = { O(t ); E1 (t ), E 2 (t ), E 3 (t ) } be an orthonormal frame fixed in the rigid body B.
We note that the point O does not necessarily belong to B.

2.3.1

Velocity Field

We recall that a vector field exists if one vector is associated with every point of E (or
part of E).

Fig. 2
Let q be a fixed point in the moving body B; it is fixed with respect to R E .
The position vector of q is expressed relative to the basis BE = ( E1 (t ), E 2 (t ), E 3 (t )) as
following:
Oq(t ) = X E1 (t ) + Y E 2 (t ) + Z E 3 (t )
where X, Y and Z are constants.
We obviously have the following derivative with respect to R e :
(

dE
dE
dE
dOq
) e (t ) = X ( 1 ) e (t ) + Y ( 2 ) e (t ) + Z ( 3 ) e (t )
dt
dt
dt
dt

and with respect to R E it is obviously:


(

dOq
) E (t ) = 0 .
dt

(0-4)

18

Chapter 0

Let us introduce the vector angular velocity Ee of R E relative to R e .


We recall that this vector does not refer to the only bases BE and Be but refers to the set of
fixed bases with respect to Be . So, since Ee does not depend on the choice of a particular
basis of R E and of R e respectively, we denote the angular velocity by (t ) or simply by .
Its well-known expression is

=(

dE
dE 2
dE
E 3 ) E1 + ( 3 E1 ) E 2 + ( 1 E 2 ) E 3
dt
dt
dt

(0-5)

with
dE i
(t ) = (t ) E i (t ) .
dt

( See e.g. Talpaert [1987] ).


Therefore, for every constant vector Op in R E , Eq. (0-4) is written:

D

dOq
) e (t ) = (t ) Oq (t ) .
dt

This vector is sometimes called the velocity vector of q relative to O.

It is the derivative of every vector fixed in the rigid body B.


Remark. The angular velocity does not depend on the choice of reference point O.
Indeed, given
(

dOq
) e = Oq ,
dt

we are going to prove that, for another point of reference O1 :


(

dO1q
) e = O1q .
dt

dO1q
) e = 1 O1q
dt

A priori, we have:

but
(

dO1q
dOO1
dOq
)e = (
)e (
) e = Oq OO1
dt
dt
dt
= (OO1 + O1q) OO1
= O1q .

Since for any O1q , we have obtained

1 O1q = O1q ,
we deduce:

1 = .

(0-6)

19

Requirements

So, we really have:


(

dO1q
) e = O1q .
dt

By taking Eq. (0-6) into account, the following equation


(

doq
doO
dOq
)e = (
)e + (
)e
dt
dt
dt

becomes:


(v q ) e = (v O ) e + qO .

(0-7)

It is the fundamental formula of rigid body kinematics.


It expresses that the velocity of every point of a rigid body B is determined if the velocity of
any point of B is known as well as the angular velocity of R E relative to R e are known.
In other words:
PR14 The velocity of any point q in R e is the sum of the velocity of O relative to R e and
the velocity vector of q relative to O.
In a natural manner, we say:
D

The velocity field of the rigid body is the mapping


(v ) e : I E E : (t , q ) a (v ) e (t , q ) = (v q ) e (t ).

This field is non uniform since it verifies the fundamental formula of rigid body kinematics.
The fundamental formula (0-7) can be found again from the formula of the vector derivative,
well-known in kinematics:
(

du
du
) e = ( ) E + Ee u
dt
dt

also denoted by
d eu d E u
=
+ Ee u
dt
dt

where u is any derivable vector function:


u : I E : t a u(t ) .

Indeed, for any points a and b of B, that is fixed in R E , we have:


(

doa
doa
)e = (
) E + Ee oa
dt
dt

dob
dob
)e = (
) E + Ee ob ,
dt
dt

and

20

Chapter 0

which implies that


d
(ob oa )) E + Ee (ob oa )
dt
= Ee ab.

(v b ) e ( v a ) e = (

Thus we have obtained:


(v b ) e = (v a ) e + ba ,
that is Eq. (0-7).
Remark. Instead of a fixed frame R e , we can consider a second moving frame:

R F = { o(t ); F1 (t ), F2 (t ), F3 (t ) }
and we say:
PR15 Given moving frames R F and R E of E, there is an angular velocity EF of R E
with respect to R F such that, for every constant vector u in R E , we have:
(

du
) F = EF u .
dt

du
)E = 0 .
dt

(0-8)

We have obviously:

In the same manner, Eq. (0-7) is generalized as it follows:


PR16 Given moving frames R E and R F of E, the velocities of fixed points a and b in
R E are connected as it follows:
(v b ) F = (v a ) F + ba EF .

(0-9)

Of course this equation as well as Eq. (0-7) recalls of course the notion of moment of a
dynam. Let us introduce this last one.
2.3.2 Dynam of Velocities
PR17 The field of velocities of points of a rigid body B is equiprojective.
Proof. Let us prove that the projections of velocities at two points of B onto the straight line
through these points are equal.
Indeed, for every pair (a, b) of points of B, we have:

ab

=c

( c R+ )

dab
d
ab = (ob oa ) . ab = 0
dt
dt

Requirements

21

v b . ab = v a . ab .

This equiprojective property means that the velocity field is really the field of moments of a
dynam.
D  The dynam of velocities or kinematic dynam of B (or R E ) with respect to R e is
defined by the following elements of reduction at q E :
- the resultant (or vector) of the dynam, which is independent of the reference point
and is denoted by (t ) ,
- the moment about q, that is the velocity v q (t ) of q such that:
v q (t ) = v O (t ) + qo (t )

where q and O are fixed in B.


We denote this dynam of velocities of B by its elements of reduction at q:


v .
q
We can generalize:
D  The dynam of velocities or kinematic dynam of a frame R E with respect to a frame
R F is defined by the following elements of reduction at a E :
- the resultant (or vector) of the dynam EF ,
- the moment about a, that is the velocity v a (t ) of a such that:
(v a ) F = (v b ) F + ab EF
where a and b are fixed in R E .
We denote this dynam by its elements of reduction at a:

EF
v ,
a
or also simply by [VEF ] , the moment of which is M a [V EF ] = (v a ) F .

2.4

ACCELERATION VECTORS
From
(v q ) e = (v p ) e + qp ,

we deduce the acceleration of q:


(a q ) e =

de
d e d e qp
(v p ) e + qp
+
,
dt
dt
dt

22

Chapter 0

but
d e qp d E qp
=
+ qp
dt
dt
= qp

and thus the fundamental formula of acceleration in rigid body kinematics is


d e
(a q ) e = (a p ) e + qp
+ ( qp)
dt

(0-10)

where the points p and q are fixed in R E .


There is another method which leads to the fields of velocities and accelerations of points of
any rigid body.
Let us consider a fixed point q of a rigid body B in motion relative to a frame
R E = { E1 (t ), E 2 (t ), E 3 (t ) }, this last one moving with respect to a frame of reference R e .
From
oq = oO + Oq ,

we deduce the absolute velocity of q, namely:


(v q ) e =

d e oO d e Oq
+
dt
dt

= (v O ) e +

d E Oq
+ Oq
dt

that is
(v q ) e = (v q ) E + (v O ) e + Oq

(0-11a)

where (v q ) E is the relative velocity of q.


Since we know that the transport velocity 1 of q is
(v q ) T = (v O ) e + Oq ,

(0-11b)

we say:
PR18 The absolute velocity is the sum of the relative velocity and the transport velocity.
In the same manner, the absolute acceleration of q is
de
de
(a q ) e =
(v q ) e =
[(v q ) E + (v O ) e + Oq ]
dt
dt

such that the terms of the sum are successively:


de
dE
(v q ) E =
(v q ) E + (v q ) E
dt
dt
1

Called Vitesse dentranement in French.

23

Requirements
de
(v O ) e = ( a O ) e ,
dt
de
d e
d E Oq
( Oq ) =
Oq + (
+ Oq ) ,
dt
dt
dt

and by denoting the relative acceleration as follows:


dE
(a q ) E =
(v q ) E ,
dt

we obtain the following expression of the absolute acceleration of q:


d e
(a q ) e = (a q ) E + [(a O ) e +
Oq + ( Oq )] + 2 (v q ) E .
dt

(0-12)

Therefore, we interpret this result as the following:


PR19 The absolute acceleration is the sum of the relative acceleration, the transport
acceleration and the Coriolis acceleration.
Example. The Euler pendulum.
Let R 0 = { o;1x0 ,1 y0 ,1z0 } be an orthonormal frame fixed in E,

R 1 = { o' ;1x1 ,1 y1 ,1z1 } be an orthonormal frame in rectilinear translation along 1x0 .


We consider a simple pendulum consisting of a mass point p suspended from the moving
point o and which is swinging in the vertical plane { o' ;1x1 ,1 y1 }. We are going to express the
velocity and acceleration

Fig. 3
Let be the angle between o'p and the downward vertical.
Let R 2 = { o' ;1x2 ,1 y2 ,1z2 ] be the orthonormal frame connected to the pendulum such that
1x2 =

o' p

o' p

24

Chapter 0

Let (v p ) i denote the velocity of p relative to R i ,

ji denote the angular velocity of R j relative to R i .


We let
oo' = X (t ) 1x0 = X (t ) 1x1 ,
o ' p = l 1x 2 .

Let us make explicit the velocity


(v p ) 0 = (v o ' ) 0 + po' 20

where p and o are fixed in R 2 .


The formula of vector derivative implies:
doo'
doo'
doo'
)0 = (
)1 + 10 oo = (
)1
dt
dt
dt
= X& 1x0 = X& 1x1 .

(v o ' ) 0 = (

In addition, the following expressions


o' p = l 1x2 = l sin 1x1 l cos 1 y1 ,
= & 1
20

z1

imply that
1x1

20 o' p = 0
l sin

1 y1
0
l cos

1z1

&
0

= l & cos 1x1 + l& 1 y1 .


Therefore, we obtain:
(v p ) 0 = ( X& + l& cos ) 1x1 + l & sin 1 y1
= ( X& + l & cos ) 1x0 + l & sin 1y0 .
We note that this result is found again directly as following:
op = oo' + o' p
= ( X& + l& cos ) 1x0 + l & sin 1 y0 ,
but
(

dop
dop
dop
)0 = (
)1 + 10 op = (
)1
dt
dt
dt

dop
) 0 = ( X& + l& cos ) 1x1 + l& sin 1 y1 .
dt

and thus

Now let us make explicit the acceleration.

Requirements

The fundamental formula of acceleration, Eq. (0-10), in rigid body kinematics is


(a p ) 0 = (a o ) 0 + po'(

d
20 ) + (20 po) 20
dt

where p and o' are fixed in R 2 .


We have:
(a o ' ) 0 = (

d
v o ' ) 0 = X&& 1x0 = X&& 1x1 .
dt

From
(

d
d
20 ) 0 = ( 20 )1 + 10 20 = && 1z1
dt
dt

we deduce:
po'(

d
20 ) 0 = l&& cos 1x1 + l&& sin 1 y1 .
dt

Furthermore, we have:
(20 po' ) 20 = l & 2 sin 1x1 + l & 2 cos 1 y1 .
We conclude:
(a p ) 0 = ( X&& + l&& cos l& 2 sin )1x1 + (l&& sin + l& 2 cos )1 y1 .
We note that we can use the method of relative motions.
Indeed, let us consider:
(v p ) 0 = ( v p )1 + (v p ) T .
But the velocity relative to R 1 is

(v p )1 = l& 1y2 = l& cos 1x1 + l& sin 1y1


and the transport velocity is
(v p ) T = (v o ' ) 0
= X& 1x0 = X& 1x1 .

( since 10 o' p = 0 )

Therefore, the absolute velocity of p is


(v p ) 0 = ( X& + l & cos ) 1x1 + l& sin 1 y1 .
As regards the acceleration vectors, we have:
(a p ) 0 = (a p )1 + (a p ) T + 210 (v p )1
= (a p )1 + (a p ) T .

25

26

Chapter 0

But
d1
(v p )1
dt
= (l&& cos l& 2 sin ) 1x1 + (l&& sin + l& 2 cos ) 1 y1

(a p )1 =

and

(a p ) T = X&& 1x1 ,

thus the absolute acceleration of p is


(a p ) 0 = ( X&& + l&& cos l& 2 sin ) 1x1 + (l&&sin + l& 2 cos ) 1 y1 .

2.5

SLIDING VELOCITY

The sliding velocity of a rigid body with respect to another one is introduced in first
courses of mechanics.
Let B1 and B2 be two rigid bodies moving in a frame of reference R.
We assume there is a point of B1 in contact with a point of B2 at each instant.
Let v (i B1 ) denote the velocity of the contact point in the motion of B1 with respect to R
at time t.
Let v (i B2 ) denote the velocity of the contact point in the motion of B2 with respect to R
at time t.

Let us recall the following


D

The sliding velocity of B2 with respect to B1 , at point i, is


v s = v (i B2 ) v (i B1 ) .

Since v (i B1 ) and v (i B2 ) have directions of the tangent plane to the bodies, at point i, we
deduce that the sliding velocity v s belongs to this tangent plane.
To conclude this introductory matter, we say:
D

Rigid bodies B1 and B2 are rolling without slipping (but pivoting) if the sliding
velocity vanishes.

27

Requirements

3. EXERCISES
Exercise 1.

Given a plane of E, we consider the set D of dynams of which the moments are
orthogonal to .
Prove that the set D of dynams orthogonal to the dynams of D is D .
Answer. Let { o; i, j } be an orthonormal frame of ,
{ o; i , j , k } be an orthonormal frame of E.
Every dynam D of D is such that:

D D , q : M q ( D) k = 0

o, q : ( M o ( D) + qo R( D)) k = 0

o, q : M o ( D) k + (qo.k ) R( D) ( R( D).k )qo = 0

o, q : M o ( D) k ( R( D).k ) qo = 0

o : [ M o ( D) // k and R( D).k = 0 ]

o : [ M o ( D) and R( D) // ] .

So, every dynam D D is of the following type:


0
a i + b j
i
j
ck = a 0 + b 0 + c k

o
o
o

(reals a, b, c) .

Let us consider the set D of dynams


i + j + k
D =

l i + m j + nk

(reals , , , l , m, n )

which are orthogonal to the ones of D ; that is such that:


0 = D. D = R( D) . M o ( D' ) + R( D' ) . M o ( D)
= (a i + b j ) . (l i + m j + n k ) + ( i + j + k ) . ck
= al + bm + c .
Since this expression vanishes for every a, b and c we necessarily have l = m = = 0 .
Therefore, every dynam D ' of D is of type:
i + j
D' =
.
nk o

We conclude that the set D of dynams of type D is really the set D of dynams of type D
whose moments are orthogonal to .

28

Chapter 0

Exercise 2.
In a frame of reference { o; i , j , k }, we consider the point a ( 2,1,1) and the sliding
vector w = i 2 j + 2k through a which defines a sliding dynam S.
(i) Determine the elements of reduction of S at o.
(ii) Find the real l such that the dynam D defined by
(l + 2) i + (l 1) j 2l k
(l 4) j + (l 4) k

is equal to S.
(iii) Prove there is another value l for which D is also a sliding dynam. Determine the axis of
the sliding dynam.
(iv) Reduce D in the case where l = 1.
Answer. (i) Since

j k

M o ( S ) = M a ( S ) + oa R( S ) = 0 + 2 1 1 = 5 j 5k ,
1 2 2

the sliding dynam S represented by


w
0
a

has elements of reduction at point o such that:


i 2 j + 2k
5 j 5k .

(ii) Since D must be equal to S, from the comparison between elements of reduction of D and
S, at o, we deduce the value l = 1.
(iii) A dynam is a sliding dynam if the scalar invariant is zero while the resultant is nonzero;
that is:
R( D) . M o ( D) = (l 1)(l 4) 2l (l 4) = 0
The solutions are:
{ l = 4 , l = 1 }.
The other value of the question is thus l = 4. In this case, the dynam is such that:
6 i + 3 j 8 k

.
0

The axis of the sliding dynam is the set of points q ( x, y, z ) such that M q = 0 , namely:

29

Requirements
i

M q = M o + qo R = 0 + x y z = 0
6

8 y + 3 z = 0

8 x + 6 z = 0
3 x + 6 y = 0,

these three equations defining the axis of the sliding dynam.


We note that this axis could be immediately found since it is the straight line through o
( M o = 0 ) and directed along R = 6 i + 3 j 8k . These equations are immediately:
x
6

y
3

z
8

(iv) For l = 1 , the dynam D is such that:


3i 2k
3 j 3 k

and the moment at every point a of the central axis is:

M a ( D) =

M o ( D) . R( D)
2

R ( D)

= 18
i 12
k.
13
13

The dynam D is the sum T = C + S of a couple and a sliding dynam such that:
0

C = 18 12
13 i 13 k

and
0

3i 2 k

3i 2k
S=
18 12 = 18
.

27
3 j 3 k o 13 i 13 k o 13 i 3 j 13 k o

Exercise 3.

A gear, represented as a disk D of fixed center o and radius R, rotates in the (x,y)-plane
of a fixed frame R e = { o;1x ,1 y ,1z } with an angular velocity .
(i) A disk D1 in the (x,y)-plane of fixed center c and radius R / 3 rolls without slipping (tooth
wheel) on the side of D and on the inside of a hoop C. Calculate the angular velocity 1 of
D1 and the angular velocity 2 of C.

(ii)

If the center of D1 is moving when D1 rolls without slipping on the side of D and on
the inside of the fixed hoop C, calculate the angular velocity 1 of D1 and the angular
velocity of oc.

30

Chapter 0

Fig. 4

Let
1u =

oc
,
oc

1v = 1z 1u .

The dynams of velocities of D, D1 and C with respect to R e are respectively:


1
VD e = z ,
0 o

1
VD1 e = 1 z ,
0 c

1
VC e = 2 z .
0 o

We specify that is positive, 1 and 2 are negative.


(i) The composition law of dynams of velocities leads to
M i (VD1 e ) = M i (VD1 D ) + M i (VD e ) .

From the end of Section 2.3, the point c being fixed when D1 moves, we have:
M i (VD1 e ) = (v c ) e + ic 1 1z = ic 1 1z .

In the same manner, the point o being fixed when D moves, we have:
M i (VD e ) = (v o ) e + io 1z = io 1z .

From these results, we deduce that the sliding velocity of D1 with respect to D, at i, is:
R

M i (VD1 D ) = 1 1v R 1v .
3

The condition of rolling without slipping is thus:

1 = 3 .
Now, let us determine the angular velocity of C. The composition law of dynams of velocities
leads to:
M j (V D1 e ) = M j (VD1 C ) + M j (VC e ) .

31

Requirements

From the end of Section 2.3, the point c being fixed when D1 moves, we have:
M j (VD1e ) = (v c ) e + jc 1 1z = jc 1 1z .

In the same manner, the point o being fixed when C moves, we have:
M j (VC e ) = (v o ) e + jo 2 1z = jo 2 1z .

From these results, we deduce that the sliding velocity of D1 with respect to C, at j, is:
M j (V D1C ) =

R
5
1 1v 2 1v .
3
3

The condition of rolling without slipping is thus:


1

2 = 1 = .
(ii) In the second situation, oc is moving and the hoop C is fixed.
Let us consider the moving frame R E = { o;1u ,1v ,1z }.
Let us calculate the sliding velocity of D1 with respect to D, at point i:
M i (VD1 D ) = M i (VD1 e ) M i (VD e ).

We have:
1
VD1 e = 1 z ,
( v c ) e c

but
(v c ) e = (
=

d oc
) e = Ee oc
dt

4
3

(since oc is constant in R E )

R 1v ,

thus we have:
1 1z
VD1 e = 4
.
3 R 1v c

From the end of Section 2.3, the point c being fixed when D1 moves, we have:
M i (V D1e ) = (v c ) e + ic D1 e = 43 R 1v R3 1 1v .
In the same manner, the point o being fixed when D moves, we have:
M i (VD e ) = (v o ) e + io = R 1v .
From these results, we deduce the following sliding velocity:
M i (VD1 D ) = ( 43 R R3 1 R ) 1v .

The condition of rolling without slipping is thus:

32

Chapter 0
4 1
3
3 1

= 0.

Now, let us calculate the sliding of D1 with respect to C, at j:


M j (VD1C ) = (v c ) C + jc D1C
= 43 R 1v + R3 1 1v .

The condition of rolling without slipping implies:


= 14 1

and finally:

1 = 32 ,

= 83 .

(since c is fixed in D1 )

CHAPTER 1

STATICS

Statics is the subdiscipline of mechanics which studies the equilibrium of material


systems under the action of forces.1
Statics is an essential branch of engineering mechanics. First, it is introduced in a classic way
by considering the resultant of forces acting on the system and the sum of moments of forces
called the total moment or simply the moment. We will quickly deal with this basic method
for which many excellent books show numerous engineering applications.
On the other hand, the method of virtual work is the subject of our full attention. This
analytical mechanics method brings into play virtual displacements or virtual velocities
(which are arbitrary vectors a priori).
In short, statics is concerned with three types of problems, namely:
-

searching for equilibrium positions of mechanical systems in presence of known forces,


looking for forces which insure the equilibrium of system,
studying the stability of equilibrium.

The classic and virtual methods have application fields which are somewhat
complementary.
So, the classic method is especially used for trying to find forces which maintain material
systems in equilibrium. For instance, this method effectively deals with obtaining necessary
and sufficient conditions to maintain a rigid body in equilibrium under the action of forces.
The method of virtual work is better suited to obtain equilibrium positions of sophisticated
mechanisms . It easily allows finding again equilibrium conditions of the classical method and
also permits the study of the stability of systems in equilibrium, and so on.
Note that the two mentioned methods can be used together in order to solve problems of
statics more easily.

A material system or mechanical system is a system of particles or a (finite) collection of rigid bodies including
possible isolated particles. For example, a mechanical system is a particle, a rigid body, connected bodies,

33

34

Chapter 1

Modeling
In theoretical mechanics, material systems as rigid bodies, mechanisms, etc. are
modeled. A modeling of a material system is not absolute but is relative to the study of the
considered system. So, various mechanical studies of a same system can lead to model this
system in different ways. For example, models for a same rigid body can have 1,2 or 3
dimensions according to the study and we say that the shapes are modeled.

1. CLASSIC METHOD
This paragraph recalls notions of first courses of mechanics.
1.1

MECHANICAL ACTIONS

1.1.1

Definitions

A cause capable of either altering the motion of a material system, maintaining it at


rest or deforming it, is called a mechanical action.
If a cause originates outside the system it is said to be an external mechanical action.
If a cause is due to an element of the material system it is said to be an internal
mechanical action.

As we shall see later, this classification can be arbitrary because of the choice of the
mechanical system isolation.

1.1.2 Forces
Forces (which are mechanical actions) exerted on a material system can be either
given forces or constraint forces, these last being generally unknown.
Given applied forces can act on a rigid body, for instance: weight force, spring force, tensile
force exerted by a flexible cable, etc. In this last case, if it is a large force compared with the
cable weight, then the action line of the tension force is the straight line of the cable; if the
weight is not negligible compared with the tension, then this last changes (in direction and
norm) along the cable. The cable exerts a force tangent to itself at the attachment point.
Besides the given forces applied to a system, there are constraint forces which constrain the
motion and also called reactions, this designation emphasizes the manner the material (to
which the body is attached) reacts. These constraint forces will later be analyzed.
Note that gravitational forces (weight force) are remote mechanical actions unlike contact
reactive forces.

1.1.3 Moment of a Force


The notion of mechanical action is generally merged with the one of force, but mixing
up these notions is justified if actions only correspond to forces.

35

Statics

However, all the mechanical actions cannot only be represented by sliding or fixed vectors
that are the forces. The reader will be persuaded by considering the classic example of a
welded connection between a rigid bar and a wall (built-in or fixed support).

Fig. 5
A force F is applied to the bar at a point x. It is intuitively obvious that the mechanical action
exerted by the bar on the wall depends on the position of x.
Therefore, in addition to F it is necessary to take into account of the moment of F about a
point a of the weld:
M a ( F ) = ax F .
Let us generalize the previous observation to collections of forces.

1.1.4 Dynam of Mechanical Action

S be mechanical systems (solid or not).


We suppose that S exerts a mechanical action on S due to several forces
Let S and

The mechanical action of

f h , h {1,,n}.

S on S is characterized by the resultant of forces:

f = fh
h

and the (total) moment about some point a:


M a = ax h f h

[ x h supp ( f h )] .1

The dynam of mechanical actions of S on S is the dynam defined by the resultant


and the (total) moment of forces exerted by S on S about a.
The resultant and moment are called the elements of reduction of the dynam at a.

supp ( f h ) is the support or line of action of

fh .

36

Chapter 1

It is denoted by
fS S
M
.
S S a
To sum up, we say:
PR1

Every mechanical action is well-defined by the dynam of the mechanical action.

In conclusion, the theory of dynams can be used in this context.


D

The dynam of mechanical actions is called the dynam of forces and is simply denoted:
f
F=
M a

(if the context is sufficiently clear).

1.2

CLASSIFICATION OF FORCES

It is necessary to specify the various types of forces which may be applied to a


material system, so we are going to give a review of these forces acting on every particle of
the system.
The important thing is to know if no external force has been forgotten or if some of them can
be neglected. In other words, the choice of external forces taken into account results from the
description of external causes acting upon the considered system. For example, for a body at
the earths surface, it is logical to neglect the influence of the moon; on the contrary, the same
does not apply to the study of tides.
The comparison between theoretical model and experience lets generally decide on choosing
the forces to consider. If the theoretical results are significantly different from observations or
expected conclusions, then supplementary forces must be eventually introduced. A famous
example was the discovery of the planet Neptune. By a careful mathematical study, Le
Verrier proved that the unexplained deviations from the predicted orbit of Uranus followed
from an unknown planet, so Neptune was located.
1.2.1 External Forces
The external forces are classified into two categories:
(i) Given external forces
They are the external forces which a priori are known at every point xh of the material system
and at each instant.
We recall that the earths gravitational force is distributed over the volume of a body or at
each point of a system. This force field exerted at each point of a rigid body may be replaced
by only one force at the center of gravity: the weight of the body.
The resultant of the given external forces acting on a mass point p h is denoted by

Fh(e) .

Statics

37

(ii) External forces of constraint

For a material system and in particular for a particle, we say:


D

A material system is subject to external constraints if there are forces originated


outside the system which restrict its motion freedom and called the external forces of
constraint.

The string of a pendulum, the walls of a container, etc. materialize external constraints.
Usually, the external constraint forces (reactions) act on isolated points of the system and
these connection forces are generally unknown and thus annoying. The method of virtual
work will allow the elimination of their effects.
In the classic example of the simple pendulum, the constraint materialized by the weightless
string prevents the suspended particle from freely moving. The string exerts a reactive force
of connection on the particle which cannot freely follow a parabola. This connection force
varies with the position and the velocity of the particle.
In the case of a frictionless connection (or smooth connection), the direction of the force is
specified.
So, the contact between a particle and a smooth surface or a smooth curve shows respectively
a connection force normal to the surface or orthogonal to the curve at the contact point.
If two bodies are in smooth contact, then the force exerted by a body on the other one is
normal to the surfaces.
We denote the resultant of the forces of constraint acting on p h by
L(eh) .
The resultant of all the external forces acting on p h is denoted by
f h( e ) = Fh(e ) + L(he) .

1.2.2 Internal Forces

The internal forces are the forces exerted between the different particles or elements of
a material system. There are two types of internal forces.
(i) Given internal forces

These forces also called the interaction forces are known forces acting between elements of
the system. At point xh , they are denoted by
Fh(i ) .
For example, they are the forces acting between the planets of the solar system (external
forces being the ones exerted by the stars of the galaxy).
A spring connecting two bodies shows another example of internal force between bodies.
(ii) Internal forces of constraint
Internal forces of constraint are (unknown) forces which make different elements of a
material system being constrained among.

38

Chapter 1

At point xh , they are denoted by


L(ih ) .
For example, an internal force of constraint is caused by a joint which constraints two bodies
to remain in contact.
The resultant of all the internal forces acting on p h is denoted:
f h(i ) = Fh(i ) + L(hi ) .
We recall that, in noninertial frames of reference, it is necessary to take supplementary forces
into account.

1.3

EQUILIBRIUM CONDITIONS

1.3.1

Definitions and Conditions

1.3.1a Definition of the Equilibrium of a Material System


D

A material system is in equilibrium relative to a frame of reference if each of its


points remains at rest relative to the frame.

This implies in particular:


D

A rigid body is in equilibrium relative to a frame of reference if its position


coordinates1 remain constant.

We can express the following (equivalent) proposition, it being understood that the velocities
of the particles are zero at a given (initial) instant.
PR2

A material system S is in equilibrium with respect to an inertial frame iff the resultant
of all the forces acting on every particle p h of S is zero:2
p h S : f h = Fh( e ) + Fh(i ) + L(he ) + L(hi ) = 0 .

(1.1)

In particular, this necessary and sufficient condition applies to rigid bodies, but obviously this
is not used in practice.
In the case of a particle, the above necessary and sufficient condition is equivalent to the
following:
D

A particle is in equilibrium (or at rest) relative to an inertial frame of reference if the


resultant of forces acting on the particle is zero.
We denote:

1
2

F +L=0.

We will see that the position coordinates will be called the generalized coordinates (distances, angles,).
The writing iff means if and only if.

39

Statics

Remark. If the frame of reference is not inertial, then it is necessary to take the forces of
transport1 and of Coriolis into account. But if we study an equilibrium with respect to the
moving coordinate system, then the Coriolis force is not to be considered since:
2 v h

(r )

=0

where v h(r ) denotes the relative velocity of p h .

1.3.1b Dynam of Internal Forces


We recall the principle of action and reaction.
PR3

The internal forces of a material system are equal in norm, opposite in direction2 and
collinear.

In other words, to every action force exerted by a particle p1 on a particle p 2 , there is a


reaction force equal in norm and opposite in a direction along the line joining the particles,
which is denoted by

f12 = f 21 .
This principle (or third law of Newton) being valid in classical mechanics, we say:
PR4

The dynam of internal forces of any mechanical system is zero.

Proof. The principle of action and reaction implies that the resultant of internal forces of a
material system S is zero:
f (i ) = f h(i ) = 0
h

because this principle is valid for every pair of particles of S.


In addition, the sum of moments of all internal forces about some point o is zero. Indeed,
successive pairs of terms such that:

ox h f h(i ) + ox k f k(i ) = (ox h ox k ) f h(i )


= xk xh K xh xk = 0

KR

imply that the (total) moment of internal forces is zero.


So, if we denote the moment of internal forces acting on any particle p h by M o ( f h(i ) ) , then
the (total) moment of internal forces of S is
M o(i ) = M o ( f h(i ) ) = 0 .
h

In conclusion, the dynam of internal forces is

1
2

Called forces dentranement in French.


Opposite in direction is said to be de sens opposs in French.

40

Chapter 1

[F ] = f
(i )

= 0.
(i )
M o
(i )

(1-2)

1.3.1c Equilibrium Conditions of a Material System

First undergraduate courses in mechanics set out the reduction, at some point, of any
collection of vectors defining a dynam D (see Requirements). Reducing a collection of (force)
vectors consists in replacing this collection with a more simple equivalent collection and it is
easily proved the following:
PR5

Every collection of (force) vectors is equivalent to a collection composed of


- a single (force) vector through an arbitrary point,
- a couple of (force) vectors.

The single vector is the resultant R(D) of the dynam and the moment of the couple is the
moment M (D) of the dynam.
The following example describes this reduction.
Example. Let us show that a collection of three forces f1 , f 2 , f 3 applied to a material
system, for instance a rigid body, is equivalent to a collection composed of the resultant of the
collection and the moment of a couple.

Let us note that here the resultant and the couple moment are not free vectors, but are bound
vectors at an arbitrary point a.
We recall that if we add two opposite vectors to a collection of vectors, then we obtain an
equivalent collection. So, every collection consisting of a vector f is equivalent to a new
collection composed of a couple and a vector f // that is the representative, at a, of the
equivalence class defined by f.
The preceding is so summed as:
{ f } { f , - f //, f // } { M, f // }
where M represents the moment of the couple.
The previous process is used with f1 , f 2 and f 3 as shown in the following figure.

Fig. 6

41

Statics

So, at a, we obtain the resultant F of three forces and the sum M a of three moments. The
collection { f1 , f 2 , f 3 } is equivalent to the collection { F , M a }.
Remark. Unlike the resultant, the moment is dependent on the choice of a. In particular, we
may choose a in order that the resultant and moment be collinear. The set of such points
defines a straight line called the central axis.

The previous proposition about the reduction of a collection of (force) vectors leads to an
interesting condition of equilibrium. Indeed, the dynam of internal forces vanishing, we can
state the following necessary condition of equilibrium:
PR6

If a material system is in equilibrium, then the dynam of external forces is necessary


zero.
This is denoted:
equilibrium
where

f ( e ) = f h(e ) ,
h

[F ]
(e)

f (e)
=
= 0.
(e)
M a

(1-3)

M a( e ) = ax h f h(e ) .
h

Remark 1. The previous condition is not sufficient to ensure the equilibrium of a material
system.
For example, we consider a plane system composed of two jointed rods. A force F is exerted
on one rod and the opposite force is exerted on the other rod.

Fig. 7

The dynam of all exterior forces is


F
F
F
F
0 + 0 = 0 + 0
p
q p
p

since F and pq are collinear.


So, this dynam is zero, but it is not sufficient for the system of rods to be in equilibrium.

42

Chapter 1

Remark 2. When studying linear momentum and angular momentum theorems, we will see
that the condition of zero exterior dynam can lead to motions of the material system (see
conservation theorems).
So, in particular, we will see that the condition F (e ) = 0 must be fulfilled for a rigid body to
be in equilibrium. But it is not sufficient to assure that the body is at rest (unless the body is
initially at rest).

[ ]

Example. Find the condition for equilibrium of a rigid body with a fixed point o, in
particular in the case of a lever.

The reaction force of constraint at o being denoted by Lo , a first condition of equilibrium (of
translation) is
F ( e) + Lo = 0
where F (e) is the resultant of given external forces.
A second condition of equilibrium (of rotation) is
M o ( F (e ) ) = 0 ,

since the constraint force does not contribute to the (total) moment at point o. This
equilibrium condition is simply denoted by
M o(e) = 0 .

Now, we consider a lever (extremities a and b, fulcrum o) which is in equilibrium under the
action of two given forces Fa and Fb .

Fig. 8

The condition of equilibrium of translation lets obtain the reaction force of constraint:
Lo = Fa Fb .

The condition of equilibrium of rotation is reduced to the following


M o( e) = oa Fa + ob Fb = 0

Fa oa sin Fb ob sin = 0

M o ( Fa ) = M o ( Fb ) .

These results are well-known.

43

Statics
1.3.2

Particular Collections of Forces Applied to a Rigid Body

A rigid body remains at rest iff the resultant and (total) moment of applied external
forces are such that
f (e ) = 0 ,

M ( e) = 0 .

The point about which the (total) moment is zero is not indicated because this moment
vanishes about any point (since M p = M q + pq f (e ) ).
Given a coordinate system oxyz, the previous necessary and sufficient conditions of
equilibrium of a rigid body are written:
f x( e) = 0 ,

f y( e) = 0 ,

f z( e) = 0

M x(e ) = 0 ,

M y(e ) = 0 ,

M z(e ) = 0 .

Now, we are going to give a review of conditions of equilibrium for particular force
collections.
(i)

Collinear forces

A rigid body is in equilibrium under collinear forces iff the resultant of forces vanishes; that
is, by choosing the x-direction of forces:
f x( e) = 0 .

In particular, if there are only two forces, these must be equal in norm, opposed in direction
and collinear.
(ii)

Coplanar forces concurrent at a point

If coplanar forces applied to a rigid body are concurrent at a point o then their moment about
o is necessarily zero.
Let oxy be the plane of forces.
The (total) moment of forces about o is
M 0( e) = 0 .

In addition, the z-components of forces are zero and thus the equilibrium conditions are
f x( e) = 0 ,

f y( e) = 0 .

In particular, a rigid body is in equilibrium under the action of three forces if the polygon of
forces is a triangle. So, three forces which ensure this equilibrium are necessarily coplanar.
Additionally, the forces must be concurrent. Indeed, if the lines of action were not concurrent,
then one of the forces would show a moment about the point of intersection of the other two,
but this would be absurd since the (total) moment of the collection of forces is zero (at any
point).
Collection of three forces plays a very important role in statics because collections of many
forces may often be reduced to this simple type.

44

(iii)

Chapter 1

Spatial forces concurrent at a point

A rigid body is in equilibrium under spatial collinear forces iff


f x( e) = 0 ,

f y( e) = 0 ,

f z( e) = 0 .

The three moment equations of equilibrium are automatically verified because the moment
about any axis through the point of concurrency vanishes.
(iv)

Coplanar and parallel forces

We consider for instance a collection of forces in a plane oxy and in x-direction.


A rigid body is in equilibrium under these forces iff
f x( e) = 0 ,
M z(e ) = 0 .

Indeed, the forces have the same x-direction and all the moments the same z-direction (normal
to the plane).
(v)

Spatial parallel forces

For example, we consider a collection of parallel forces in the x-direction.


A rigid body is in equilibrium under these forces iff
f x( e) = 0 ,
M y(e ) = 0 ,

M z(e ) = 0 .

Indeed, the force components in y and z are zero and the (total) moment of forces has no
component in x.
(vi)

Forces concurrent with a line

For example, we choose this line as x-axis.


The moment of any (force) vector about an axis is zero if the line of action and axis are
concurrent. So, the collection of forces automatically verifies the condition M x(e ) = 0 and
therefore the conditions of equilibrium are

(vii)

f x( e) = 0 ,

f y( e) = 0 ,

M y(e ) = 0 ,

M z(e ) = 0 .

f z( e) = 0

Coplanar forces

For example, the forces are assumed in the plane oxy.


The conditions of equilibrium of a collection of such forces are immediately:
f x( e) = 0 ,
M z(e ) = 0 .

f y( e) = 0 ,

Statics
1.4

45

TYPES OF EQUILIBRIUM

The necessary and sufficient conditions of equilibrium of a rigid body under a


collection of forces let calculate (at equilibrium) the unknown forces of reaction, but these
conditions may not be sufficient to evaluate all of them. This is function of types of
constraints.
D

A collection of forces is hyperstatic if the number of unknown forces is larger than the
number of equations of equilibrium condition.
A collection of forces is hypostatic if the number of unknown forces is smaller than
the number of equations of equilibrium condition.
A collection of forces is isostatic if the number of unknown forces is equal to the one
of equations of equilibrium condition.

Remark. In the hypostatic case, where the body is called partially constrained, the
equilibrium is instable since the unknown forces cannot satisfy all the equilibrium equations.
More explicitly, for each unsatisfied scalar equation of equilibrium, there is a corresponding
possible motion, either of translation (for an equation of force), or of rotation (for an equation
of moment).

In engineering mechanics, it is essential to know the forces acting at different points of a solid
structure. These forces can deform and break the structures; but often and in first
approximation, these structures can be considered as rigid bodies. Courses of strength of
materials detail these notions.
D

A structure is statically determinate if there is the minimum number of constraints


necessary to ensure the equilibrium.

In this case, from given external forces, all the unknown forces and moments acting on each
part of the structure can be determined.
D

A structure is statically indeterminate if there are more constraints than necessary to


maintain an equilibrium.
In other words:
if there are more unknown components of reaction forces than scalar equilibrium
equations.

So, a statically indeterminate structure is such that the forces acting on its parts are not
completely determined by the given (external) forces, owing to stresses within the structure.
This type of problem is more difficult to solve than the one of statically determinate
structures.
D

Constraints are said to be redundant if they can be removed without destroying the
equilibrium.

Example 1. A rigid body fixed at two points o1 and o2 is generally a matter for hyperstatic
equilibrium.

46

Chapter 1

Let d be the distance between these points,


{ o1 ; i , j , k } be an orthonormal frame of reference where k is collinear to o1o 2 and has the
direction1 of o1o2 ,
L1 ( L1x , L1 y , L1z ) and L2 ( L2 x , L2 y , L2 z ) be the unknown constraint forces of reaction at

respective points o1 and o2 ,


F ( Fx , F y , Fz ) be the resultant of given external forces acting on the body.
The condition of (translation) equilibrium
F + L1 + L2 = 0

is equivalent to the system of 3 scalar equations:


Fx + L1x + L2 x = 0 ,
F y + L1 y + L2 y = 0 ,
Fz + L1z + L2 z = 0
which contains 6 unknowns.
We are going to express the condition of rotation equilibrium, for instance at o1 .
Since the (total) moment of forces about o1 is composed of:
-

the (total) moment of given external forces M o(1e) ( M x , M y , M z ) ,

the moment of the constraint force of reaction L2 , namely:


M o1 ( L2 ) = o1o 2 L2 = d L2 y i + d L2 x j ,

then the conditions of rotation equilibrium are written:


M x d L2 y = 0 ,

M y + d L2 x = 0 ,

Mz = 0.

Be careful! This last equation imposes one condition, namely: the (total) moment of external
forces about the fixed line o1o2 must be zero. The remaining conditions of equilibrium allow
determining the components of constraint forces:
L2 x = M y d ,

L2 y = M x d .

Next, the first two equations of translation equilibrium allow obtaining L1x and L1 y .
So, there is only one equation at our disposal:
L1z + L2 z = Fz ,
but we must calculate two unknowns L1z and L2 z !
The collection of forces is hyperstatic since there are more unknowns than equations. It is
possible to make it isostatic by canceling, for instance, the component L2 z with the aid of a
sliding guide supporting only forces normal to the guide.
Example 2. The collection of forces which fix a rigid body at three points is hyperstatic in
general.

Indeed, there are 3 constraint (reaction) forces, that is 9 unknowns for 6 equilibrium equations.
1

Sens dorientation in French

Statics
1.5

47

STRESS AND CONTACT DYNAM

In reality, forces between bodies in contact are distributed over an area; that is,
concentrated forces do not exist, but are distributed over a region: there is a force field! This
remark also holds in the cases of lines and volumes.
Let B1 and B2 be rigid bodies in contact,
S be the contact surface,
be the tangent plane to bodies at any point p S .
The body B1 exerts a (contact) mechanical action on B2 at every point of S.

Fig. 9
Remark. We can also imagine any surface splitting two parts of a body.

1.5.1 Stress
D

The stress of B1 on B2 , at p, is the force exerted by B1 on B2 per unit area.

It is denoted by
l (p12)

or simply by
l

(if no possible confusion).

The usual units for stress are:


- the pascal or newton per square metre: Pa = N / m 2 ,
the megapascal: MPa = 10 6 Pa = N / mm 2 = 10 bars .

The vector component of stress perpendicular to the tangent plane is called the
normal stress and the vector component of stress tangent to the plane is called the
shear stress.1

If we denote the shear stress by t and the normal stress by n, at any p , then we express
the corresponding stress by
l = t + n.
1

Unlike the rigid bodies for which shear stresses exist, the fluids at rest show only normal stresses.

(1-4)

48

Chapter 1

Fig. 10
By considering a surface splitting two parts of a body, we also define:
D

A compression (resp. a tension) is a stress perpendicular to the surface which pushes


(resp. pulls) the material of each part towards the other.

The next figure shows various stresses between materials of part 1 and part 2 in the cases of
compression ( C ), tension ( T ) and shear stress ( SS ).

Fig. 11
1.5.2 Contact Dynam

Let us define the dynam of the mechanical action of contact of B1 on B2 , simply


called the contact dynam of B1 on B2 or the constraint dynam of B1 on B2 .
Let o be a reference point.
D

The contact dynam of B1 on B2 is defined by its elements of reduction, at o, namely:


L(12) = l dS

(1-5)

M 0(12) = op l dS .

(1-6)

It is denoted by
L(12)
(12) .
M
o

Statics

49

Remark 1. The contact dynam is obtained from the stress l of B1 on B2 known at every
point of S.
Remark 2. We denote by L(12) the resultant of forces acting on an element of S (at p) and
by A the corresponding area.
L(12)
The resultant per unit area is obviously

The stress acting on the surface element is the vector defined by


L(12)

A0
A

l (12) = lim

The situation is less simple than the one of remark 1. The stress l at p S depends upon the
choice of particular surface element.
Remark 3. We must distinguish body forces from surface forces.

Body forces are distributed forces applied over the matter space occupied by the body.
An obvious example is given by the gravitational attraction. The acceleration due to gravity
being g ( m / s 2 ) and the mass density of the body being ( kg / m 3 ), then the corresponding

body force per unit volume is g ( N / m 3 ).


The terminology body force density is sometimes used for the preceding.
Surface forces are forces acting on the elements of a surface region of a body.
For example, stresses show such distributed forces over the contacting surfaces of two bodies
or pressure exerted by fluids, etc. They are forces per unit area.
The terminology surface force density is sometimes used for the preceding.

Now, we introduce the notions of (total) tangential force and (total) normal force for a contact
surface S.
D

The (total) tangential force, denoted by T (12) , is the tangential vector component of
the resultant of the contact dynam.
The (total) normal force, denoted by N (12) , is the normal vector component of the
resultant of the contact dynam.

In other words, we denote:


T (12) = t dS ,

(1-7)

N (12) = n dS .

(1-8)

The rolling moment, about p, denoted M tp , is the tangential part of the moment of
contact dynam.
The pivoting moment, about p, denoted M np , is the normal part of the moment of
contact dynam.

50

Chapter 1

To sum up, we simply denote:


L=T + N ,

(1-9)

M p = M tp + M np .

(1-10)

Remark. The contact between two bodies strictly at a point is an idealized situation; in
reality, there is always a contact area. If such an ideal contact is thinkable, then we naturally
assume that the moment of contact dynam vanishes at the contact point.

1.5.3 Dry Friction and Coulombs Laws

Perfectly smooth contact between two bodies is evidently an ideal situation with only
normal forces of reaction. In most of cases, additional forces must be taken into account to
describe the real process.
D

Friction forces are forces tangent to the plane of contact between bodies and opposing
the (impending) motion of one surface relative to the other.

Friction forces are classified according to the nature of contacting surfaces, there are dry,
fluid, internal frictions for instance.
In particular, a dry friction can be roughly considered as an effect resulting from microscopic
irregularities, from molecular attraction, etc. present in (rough) surfaces of two solids in
contact.
This type of friction is sometimes called Coulomb friction because, in 1781, Charles de
Coulomb developed elementary laws describing dry frictions.

1.5.3a Friction Force and a Classic Experiment

Before showing the Coulombs laws, we consider the following classic experiment.
A solid block B2 of mass m is in rest on a horizontal plane B1 . The contacting surfaces show
a certain extent of roughness.

Fig. 12

The experiment consists in exerting on B2 a given horizontal force F with an increasing


magnitude in order to set the block in motion.

Statics

51

In this problem with friction, the reaction force L of B1 on B2 is the resultant of two forces
exerted by the plane on the block:
- a (tangential) friction force, denoted T, which is in a direction to oppose motion (or
motion tendency),
- a (normal) friction force, denoted N, which balances the weight mg.
By enlarging details of the contact surfaces, we would see irregularities.

Fig. 13

At each hump there is a reaction force L(i ) = T(i ) + N (i ) . The (total) tangential friction force T
is the sum of various T(i ) .
First, we consider the block at rest.
In this case of equilibrium, we have in particular:
F =0

T =0.

The reaction force L is normal to B1


L=N

and it balances mg.


As soon as F is different from zero, but the block being motionless, we have:
T = F .

In this case, T is called the static friction force and is denoted by Ts .


There is a maximum value of Ts which corresponds to the maximum magnitude of F for
which B2 remains at rest. It is denoted by
Tsmax .

The situation of static friction is summed as follows:


T = F ,

0 Ts Tsmax .

Second, for a sufficient force F the slipping motion of the block is possible. As soon as this
motion starts then T is called the kinetic friction force and is denoted by Tk .

52

Chapter 1

The following graph shows both situations.

Fig. 14

Along the ordinate of this graph we plot the norm of the tangential force of friction and along
the abscissa we plot the norm of the given force F.
As soon as the sliding begins, the friction force magnitude falls for a very short time and next
decreases slightly.
So, there are two regions, the one of the static friction force and the other one of the kinetic
friction force beginning with the drop of the friction force. A drop explanation is possible by
considering the irregularities of contact forces [see e.g. Meriam (1975)]. If the surfaces are in
relative motion, then the contacts occur near the tops of the humps where the action lines of
various reaction forces L(i ) imply that the corresponding tangential friction forces T(i ) are
smaller than when the surfaces are in relative rest. This can explain that the force F for
maintaining the block motion is smaller than the one to set the block in motion.
Remark. In the case of surfaces B1 and B2 sliding over each other, the sliding velocity of
B2 relating to B1 at some point x belonging to the tangent plane is denoted by v; then the
condition expressing that the friction force T must be in a direction opposite to the direction
of relative motion is expressed as:
T v =0

and

T .v < 0 .

1.5.3b Coulombs Laws

We know that if the constraints are not smooth, besides the normal stresses
supplementary unknowns must be taken into account in problems of frictional resistance;
namely: the shear stresses. Therefore, it is necessary to introduce new relationships called the
laws of friction. These laws are different according to the problem is concerned by
equilibrium (statics) or not.
Let us show the friction laws of Coulomb.

53

Statics

At rest, just before the impending motion, it is observed that the maximum value of the static
friction force is proportional to the normal force; that is:
Tsmax = f s N

(1-11)

where the (positive) coefficient of proportionality f s is called the coefficient of static friction.
To conclude the case of no relative motion of surfaces, we write the following general
condition:
Ts f s N
(1-12)
where the equality sign corresponds to the impending motion.
In the case of relative motion of contact surfaces, a simple law is also observed: The norms of
the kinetic friction force and normal force are proportional
Tk = f k N

(1-13)

and the coefficient of proportionality f k is called the coefficient of kinetic friction.


Remark 1. Roughly speaking, coefficients of friction greatly depend on the type of materials
in contact. For instance:

Material
Steel on steel (dry)
Steel on steel (greasy)
Brake material on cast iron
Rubber on wood
Mild steel on mild steel
Aluminum on mild steel
Teflon on steel
Cast iron on cast iron

fs

fk

0.6
0.1
0.4
0.4
0.74
0.61
0.04
1.1

0.4
0.05
0.3
0.3
0.57
0.47
0.04
0.15

All these values depend on the experiment conditions (lubrification, polishing,


temperature,) and the last example shows a coefficient higher than unity.
Remark 2. In the majority of cases, the coefficient of kinetic friction is a lower value than
the coefficient of static friction:

Tk Tsmax

fk fs .

To conclude this section, we consider very interesting angles.


When two surfaces are in contact, experiments (and intuition) point up that the angle
between L and N is larger for rough surfaces than for smoother surfaces. It is quite natural to
consider the tangent of this angle:
tan =

T
N

(1-14)

54

Chapter 1

So, for the maximum value of the static friction force, there is a specific angle s such that
tan s = f s ,

(1-15)

where the angle s is called the angle of static friction.


In the same manner, in the case of relative motion, we consider the angle k defined by

tan k =

Tk
N

= fk

(1-16)

which is called the angle of kinetic friction.


Of course, we have:

s .
So, there is a limiting position of the reaction force L; a right circular cone of vertex angle
2 s is defined and said to be the cone of static friction, the line of action of possible reaction
force L, at impending motion, being a generating line of the cone boundary.

Fig. 15
When motion occurs, the reaction force line lies to the surface of a cone of vertex angle 2 k
called the cone of kinetic friction.

1.6

TYPES OF CONSTRAINTS

Let B1 and B2 be two rigid bodies in contact,


S be the portion of contact surface,
be the tangent plane to B1 and B2 at any point p S .
In the following study of particular constraints, we will only consider constraints without
friction.

55

Statics

We recall the following.


D

A constraint between B1 and B2 is frictionless or smooth if the stress is such that


p S : l p .

In theoretical mechanics, the constraints between rigid bodies (in relative motion) are
modeled.
Let { o; i , j , k } be an orthonormal frame of reference such that the origin belongs to each
constraint, but we emphasize this frame is not fixed in connected bodies.
If the constraint dynam of B1 on B2 is defined by
L(12) = X i + Y j + Z k ,
M 0(12) = L i + M j + N k

(1-17)

then we denote this contact dynam by


X Y Z
L M N .

In order to make clear the various types of constraints, it is necessary to define the following.
D

The number of degrees of freedom of a constraint is the minimum number of possible


independent motions of translation and rotation, relating to x, y and z-axes, of B2 with
respect to B1 .

Let us introduce different smooth constraints.


1.6.1 Punctual Constraint
D

Rigid bodies B1 and B2 are said to be constrained at a point or to be connected by a


punctual constraint if, during their relative motion, a point of one body remains in a
plane fixed in the other one.

For instance, two cylinders such that their respective generating lines are not parallel show a
punctual constraint.
Let us search for the number of degrees of freedom.
Let q be a point of B2 remaining in a plane of B1 .
Let oxyz be a coordinate system such that the contact point q is at point o and z-axis is normal
to the plane .
The allowed motions of B2 relative to B1 are of 5 types, namely:
x and y-translations,
rotations about ox, oy and oz,
since in these cases, the point q definitely remains in .
So, we say:

56

PR7

Chapter 1

The number of degrees of freedom of a punctual constraint is 5.

Fig. 16
The hypothesis of a strictly punctual constraint implies that the moment of the contact dynam
is zero. In addition, the hypothesis of a smooth connection means that the resultant L(12) of
the contact dynam is normal to (since l o ).
So, we state:
PR8

A punctual constraint is characterized by a contact dynam of the following type


0 0 Z
0 0 0

and is schematized as follows:

Fig. 17
1.6.2 Rectilinear Constraint
D

Rigid bodies B1 and B2 are connected by a rectilinear constraint if, during their
relative motion, a straight line D of one body remains in a plane fixed in the other
one.

Two cylinders with parallel generating lines, a cone connected with a plane, etc. are examples
of rectilinear constraints.
Let oxyz be a coordinate system such that o belongs to the straight line D of B2 , x-axis is
collinear to D and z-axis is perpendicular to .

Statics

57

Fig. 18
The allowed motions of B2 relative to B1 are of 4 types, namely:
x and y-translations,
rotations about ox and oz,
since in these cases D definitely remains in .
So, we say:
PR9

The number of degrees of freedom of a rectilinear constraint is 4.

The hypothesis of a smooth connection materialized by the x-axis means that the resultant of
the contact dynam
L(12) = l p dx
ox

is along oz.
In addition, the moment of contact dynam about o

M o(12) = op l p dx
ox

is collinear to the y-axis.


So, we have:
PR10 A rectilinear constraint shows a contact dynam of the following type
0 0 Z
0 M 0

and is schematized as follows:

Fig. 19

58

Chapter 1

1.6.3

Annular-Linear Constraint

Rigid bodies B1 and B2 are connected by an annular-linear constraint if, during their
relative motion, a point q of one body remains on a straight line D fixed in the other
body.

A spherical surface tangent to a cylindrical surface in contact along some great circle is an
example of such a constraint.
The straight line D is perpendicular to the plane of this circle.

Fig. 20
Let oxyz be a coordinate system such that the x-axis is collinear to the straight line D of B1 ,
the point q B2 being the origin (at the center of the sphere).
The allowed motions of B2 relative to B1 are of 4 types, namely:
x-translation,
rotations about ox, oy and oz,
since in these cases q definitely remains on D.
So, we say:
PR11 The number of degrees of freedom of an annular-linear constraint is 4.
The hypothesis of a smooth constraint means that the line of action of the stress l p is through
the origin o and perpendicular to the x-axis at each point p of the connection.
Therefore, the resultant L(12) of the contact dynam belongs to the plane oyz and the moment
of this dynam is zero.
So we say:
PR12 An annular-linear constraint shows a contact dynam of the following type:
0 Y
0 0

Z
0 o

and is schematized as follows:

Statics

59

Fig. 21

1.6.4 Ball-and-Socket Joint


D

Rigid bodies B1 and B2 are connected by a ball-and-socket joint if, during their
relative motion, a point q connected to one body remains at a point r fixed in the
other one.

Two concentric spherical surfaces of same radius exemplify such a constraint.


Let oxyz be a coordinate system with origin at q r .

Fig. 22
The allowed motions of B2 relative to B1 are of 3 types:
rotations about x, y and z-axes,
since in these cases q and r coincide.
So, we say:
D

The number of degrees of freedom of a ball-and-socket joint is 3.

The hypothesis of a smooth constraint means that the line of action of the stress l p is trough o
at each point p of the connection.

60

Chapter 1

So, we obtain:
PR14 A ball-and-socket joint shows a contact dynam of the following type
X
0

Z
0 0 o

and is schematized as follows:

Fig. 23
1.6.5 Plane Support
D

Rigid bodies B1 and B2 are connected by a plane support if, during their relative
motion, a plane connected to one body coincides with a plane connected to the other
one.

Examples.

Rocker, roller, ball, supports

Fig. 24
We consider two (coinciding) plane surfaces of connection.
Let oxyz be a coordinate system such that o belongs to the common plane and z-axis
perpendicular to this plane.
The allowed motions of B2 relative to B1 are of 3 types, namely:
x and y-translations,
rotation about oz,
since in these cases the coincidence of planes connected to B1 and B2 respectively is
preserved.
So, we get:
PR15 The number of degrees of freedom of a plane support is 3.

Statics

61

Fig. 25
The hypothesis of a smooth constraint means that the line of action of the stress l p is
collinear to the z-axis at each point p of the connection.
Therefore the resultant L(12) of the contact dynam is normal to the supporting plane surface
and the moment of this dynam is tangent to the plane oxy.
So, we say:
PR16 A plane support shows a contact dynam of the following type
0 0 Z
L M 0

and is schematized as follows:

Fig. 26
1.6.6 Sliding Pivot
D

Rigid bodies B1 and B2 are connected by a sliding pivot or a sliding hinge if, during
their relative motion, a straight line fixed in one body coincides (but slides!) with a
straight line fixed in the other body.

The tangent surfaces of connection are revolution cylindrical surfaces.


Example. A collar free to move smoothly along a rod.

62

Chapter 1

Let oxyz be coordinate system such that x-axis and cylinder axis coincide.

Fig. 27
The allowed motions of B2 relative to B1 are of 2 types:
x-translation,
rotation about ox,
since in these cases the mentioned straight lines coincide.
Thus we say:
PR17 The number of degrees of freedom of a sliding pivot is 2.
The hypothesis of a smooth guide means that the line of action of the stress l p (the resultant
of action dynam as well) is parallel to the plane oyz and is through the x-axis, at each point p
of the contacting surface.
The moment of l p about the x-axis is zero and thus the moment of contact dynam has no xcomponent.
Thus we have:
PR18 A sliding pivot shows a contact dynam of the following type
0 Y
0 M

Z
N o

and is schematized as follows:

Fig. 28

Statics

63

1.6.7

Sliding Guide

Rigid bodies B1 and B2 are connected by a sliding guide if, during their relative
motion, a plane fixed in one body coincides with a plane fixed in the other one
and if a straight line of the first plane coincides with a straight line of the other plane.

The surfaces of connection are ruled (but not of revolution) with generating lines parallel to a
common straight line D.
Example.

Fig. 29
Let oxyz be a coordinate system such that the x-axis is along the common straight line D.

Fig. 30
The allowed motion of B2 relative to B1 is only of type:
x-translation,
since in this case the mentioned straight lines and planes of B1 and B2 coincide respectively.
Thus we can state:
PR19 The number of degrees of freedom of a sliding guide is 1.
The hypothesis of a smooth connection means that the line of action of the stress l p (the
resultant of action dynam as well) is parallel to the plane oyz. But, unlike the sliding pivot,
this line of action is not (necessarily) through the x-axis and the moment of contact dynam has
a component in the x-direction.

64

Chapter 1

So, we get:
PR20 A sliding guide shows a contact dynam of the following type
0 Y
L M

Z
N o

and is schematized as follows:

Fig. 31
1.6.8 Screw Joint
D

Rigid bodies B1 and B2 are connected by a screw joint if, during their relative motion,
a straight line of one body, e.g. B2 , coincides with a straight line of the other body B1
called the helicoidal axis D of a circular helix bound to B1 . In addition, a point
connected to B2 follows the previous circular helix.

The surfaces of connection are helicoid surfaces of which the points, in relative motion,
follow helixes.
Examples of such a connection are well-known (jack,).
Let oxyz be a coordinate system such that x-axis is along the straight line D.

Fig. 32
The allowed motion of B2 relative to B1 is of type:
x-translation combined with a rotation about x-axis.
The rotation is not independent of the translation because there is an obvious equality:

Statics

x=

h
2

65

where h is the pitch of the helix and the rotation angle.


Consequently, we have:
PR21 The number of degrees of freedom of a screw joint is 1.
The inclination angle of the circular helix is defined by
tan =

(=

2 R

x
)
R

where R is the circle radius.


The relationship between translation and rotation is expressed as
x = R tan .

By considering a smooth connection, from the expression of l p it can be proved the following
M o(12) . i =

h
2

i . L(12) .

Thus we express:
PR22 A screw joint shows a contact dynam of the following type

Y
X
h

X M
2

N
o

and is schematized as follows:

Fig. 33
1.6.9 Pivot
D

Rigid bodies B1 and B2 are connected by a pivot if, during their relative motion, two
different points of one body continuously coincide respectively with two points
fixed in the other body.

It is the most frequent connection. For example, all the mechanisms which transform motions
of rotation have pivots.

66

Chapter 1

The following pin connection is an example of pivot.

Fig. 34
The surfaces of pivot connection are non cylindrical surfaces of revolution.
Let a2 and b2 be points of B2 that respectively coincide with a1 and b1 of B1 .
Let oxyz be a coordinate system such that the z-axis is the rotation axis (a1b1 ) (a 2 b2 ) .

Fig. 35

The allowed motion of B2 relative to B1 are only rotations about oz.


So, we say:
PR23 The number of degrees of freedom of a pivot is 1.
The hypothesis of a smooth connection means that the line of action of the stress l p (the
resultant of action dynam as well) is through the z-axis and thus the moment of l p about this
axis is zero.
So, we express:
PR24 A pivot shows a contact dynam of the following type
X Y Z
L M 0

and is schematized as follows:

Fig. 36

Statics

67

1.6.10 Embedding or Welded Joint


D

Two rigid bodies B1 and B2 are welded or embedded if, during their relative motion,
three different points of one body continuously coincide respectively with three points
fixed in the other body.

No relative motion is allowed. The mechanical actions on one body are transmitted to the
other body.

Fig. 37

Thus we have:
PR25 The number of degrees of freedom of an embedding is zero.
PR26 An embedding shows a contact dynam of the following type
X Y Z
L M N .

A welded connection shows the greatest degree of fixity that a connection can impose to a
body. It is generally not schematized.
To conclude, we give the plane representations of the essential previous connections.
1. Annular linear constraint

2. Ball-and-socket joint

68

Chapter 1

3. Plane support

4. Sliding pivot

5. Sliding guide

6. Screw joint

7. Pivot

Fig. 38
1.7 FREE-BODY DIAGRAM

In classical methods of statics, it is essential to determine all the forces required to


prevent a body or a system of bodies from moving. In equilibrium studies, any action cannot
be forgotten.
In order to take account of all given constraint forces, we must introduce the important
method of isolation of body or system of bodies to consider. This so-called free body diagram
method consists in choosing a body or bodies to be isolated from surroundings.
The isolation choice is a very important stage of the method!
Next, once a decision is taken about which bodies are to be considered then these connected
bodies are viewed as a single body enclosed by a boundary materialized by the isolation.
Finally, all the forces acting on the isolated body (that is, the forces applied by the removed
contacting and attracting bodies) must be represented.

Statics

69

Example. Free-body diagram of a plane frame.

We consider the following structural elements that altogether form a rigid framework, that is
considered as a single rigid body.

Fig. 39

The pivot at point a exerts, on the frame, a reaction force L with


- a vertical component that must be directed up to be opposed to weights W1 and W 2 ,
- a horizontal component that must be directed to the right to be opposed to the tension T.

Fig. 40

It is advisable to show the chosen coordinate axes.


Remark 1. In 3-dimensional problems, the free-body diagram can be shown in various
coordinate planes.
Remark 2. If the direction of a constraint force is unknown, then it is arbitrarily chosen, but
the right direction is later deduced from calculus.
Example. Draw the free-body diagram for a horizontal pole of mass m maintained in
equilibrium under the action of two forces exerted by two cables fixed at points b and c of the
pole; a ball-and-socket joint prevents the motion at point a.

Fig. 41

70

Chapter 1

In the free-body diagram, the x- and z-directions of La are arbitrarily chosen, but the ydirection is certainly correct.
Remark 3. In the study of systems of interconnected rigid bodies, it is necessary to draw a
free-body diagram for each body if we want to bring the internal forces of constraint into
prominence.
It is essential to make an appraisal of all the forces applied to each solid and it is highly
advisable to be careful of various types of constraints present in the system.

The choice of the directional sense of unknown forces of constraint is not arbitrary since it
must be consistent with the connections of rigid bodies.
For example, we consider two rods pinned at both of their ends.

Fig. 42

2. METHOD OF VIRTUAL WORK


One of the essential goals of the virtual work method is to establish the equilibrium
positions of a system of particles (or of a particle) by introducing arbitrary vectors a priori;
such vectors will next be called either virtual displacements 1 or virtual velocities according to
the considered context.
The method of virtual work lets determining the unknown forces that are necessary and
sufficient to maintain the equilibrium state of systems, it permits to know a lot about stability
of systems in equilibrium too.
In many studies, the virtual work method leads to faster results than the classical method,
because its main advantage consists in eliminating the part of unknown reaction forces.
Both methods contribute to the understanding of the mechanics of systems, each having its
more or less suitable application domain; however their simultaneous use can be profitable.
An essential advantage of the virtual work method lies in the fact that the system as a whole is
viewed, there is no dismemberment of the system and the analysis of the forces that must be
considered is more reduced. So, in the case of interconnected systems of bodies for instance,
the method of virtual work will be more advisable.

This concept was introduced by Jean Bernoulli in the 18th century (under another formulation obviously).

71

Statics

2.1

NUMBER OF DEGREES OF FREDOM GENERALIZED COORDINATES

First, we recall the following definition.


D

A system of particles (a particle in particular) is subject to constraints if it cannot


freely move.
In other words:
if the motion of particles is constrained.

Constraints can be specified in different ways, either by a restriction on the coordinates, or on


the velocities or both of them.
Later, we will meet several types of constraints.

2.1.1 Number of Degrees of Freedom


D

The number of degrees of freedom of a mechanical system is the minimum number of


coordinates1 (or parameters) necessary to describe the motion of the particles of the
system.

It is denoted by n.
Example 1. We consider a particle moving on a surface whose equation (of constraint) is the
following relation between Cartesian coordinates and time:
f ( x, y , z ; t ) = 0 .

The particle has two degrees of freedom since two independent coordinates are required to
specify the position of the particle, at any instant, on the (possibly moving) surface.
In other words, two parameters are necessary and sufficient to determine the particle position
at any instant.
Example 2.
constraint):

A particle moving along a curve defined by the following equations (of


f ( x, y , z ; t ) = 0

g ( x , y , z; t ) = 0

has one degree of freedom.


Only one parameter is necessary and sufficient to specify, at any instant, the position of the
particle on the (possibly moving) curve.
Example 3. Evidently, a particle free to move in space has three degrees of freedom, since
three independent coordinates are necessary and sufficient to specify the position of the
particle at any instant.

So, in a general way, the number of degrees of freedom of a particle in space is


n = 3m
1

That is: Independent coordinates

(0 m 2)

72

Chapter 1

where m is the number of constraint equations of type:


f ( x, y , z ; t ) = 0 .

Example 4. A rod supported by two springs has two degrees of freedom, since each spring
which is assumed to oscillate in a vertical line shows one degree of freedom.

In the first two examples of a particle subject to one or two constraints, the Cartesian
coordinates of the particle are not independent and thus are overabundant coordinates. We
say:
D

Overabundant coordinates are called primitive coordinates.

They are denoted by u j .

2.1.2 Generalized Coordinates


D

The generalized coordinates 1 are the independent coordinates required to specify the
position of each particle of a system.

The generalized coordinates let describe the motion of a mechanical system. They are as
many as degrees of freedom.
Generalized coordinates are denoted by
qj

( j = 1,..., n ).

Remark 1. Generalized coordinates can just as well represent distance coordinates as angles;
we will see examples of this.
Remark 2. The choice of generalized coordinates can be multiple; however these coordinates
will be chosen so as to simplify calculations.
D

The space of generalized coordinates is called the configuration space.

Every configuration space of a mechanical system with n degrees of freedom has n


dimensions.
A point (q1 ,..., q n ) of the configuration space determines the position of the mechanical
system at a given instant; in other words this (descriptive) point gives the configuration of the
system.
The evolution of the mechanical system in the course of time makes describe to (q1 ,..., q n ) a
curve in the configuration space and we say:

Also called Lagrange coordinates.

73

Statics

The generalized trajectory of a mechanical system is the curve followed by the point
(q1 ,..., q n ) in the configuration space.

The generalized velocity associated with the generalized coordinate q k is the time
derivative q k .

For example, the generalized velocity associated with an angle is the angular velocity  .
Example 5. Given a free particle moving in 3-dimensional space, the three chosen
generalized coordinates are either the Cartesian coordinates, or the three spherical coordinates
or the three cylindrical coordinates, the choice depending on the geometrical context
(configuration!).

The configuration space is the 3-dimensional space R 3 .


Example 6. A particle moving along an ellipse defined by
x = a cos ,

y = b sin ,

z = 0,

( a, b R )

has one degree of freedom, the generalized coordinate being the polar angle .
The configuration space is one-dimensional.
Example 7. A particle moving on a spherical surface has two degrees of freedom. So is the
pendulum bob constrained to move on a spherical surface of radius R.

The bob is free to swing through the entire solid angle about a point, it is located by two
generalized coordinates which are the following spherical coordinates: the colatitude and
the longitude .
The overabundant Cartesian coordinates are related to generalized coordinates as follows:
x = R sin cos ,

y = R sin sin ,

z = R cos .

The configuration space is a 2-dimensional space: the sphere S 2 .


Example 8. A plane double pendulum is an example of mechanical system which has two
degrees of freedom.

Generalized coordinates are the respective angles that measure the deviations of each of two
weightless rigid rods from the vertical.
The configuration space is a 2-dimensional space: the torus T 2 = S 1 S 1 , Cartesian product
of two circles.
Example 9. A system of two particles linked together by a rigid weightless rod and moving
in space has five degrees of freedom.

Generalized coordinates are the three coordinates of the center of mass of the system and two
angles determining the inclination of the rod in space.
The configuration space is a 5-dimensional space.

74

Chapter 1

In another way, a system of two particles in space has 3 N = 6 degrees of freedom a priori, but
there is a constraint expressing that the distance between particles remains constant. This
reduces to 5 the number of degrees of freedom.
This last example suggests that constraints reduce the number of degrees of freedom.
In a general way, a set of N particles in R 3 subject to constraints defined by m equations
f j ( x, y, z ) = 0 has
n = 3N m
degrees of freedom.
Exercise 10. A rigid body B is an obvious example of system of particles subject to
constraints, since the mutual distances of all pairs of particles remain constant, that is:

a, b B : ab = ( xb x a ) 2 + ( yb y a ) 2 + ( z b z a ) 2 = k

( R+ ).

It would be impossible to exactly describe the motion of any system of N ( 3 ) points if it


was not a rigid body.
In fact, it is sufficient to know the positions of three particles of a rigid body to know the ones
of all particles of the solid. The rigid body would have nine (3 3 ) degrees of freedom a
priori.
But there are three constraints that are the three mutual distances ( 12 N ( N 1) = 3 ).
Therefore, every position of a free solid is completely determined by six parameters or
generalized coordinates ( n = 9 3 = 6 ).

We can choose as generalized coordinates: the three coordinates of the center of mass and
three parameters describing the body inclination for instance by the three Euler angles.
Example 11. What is the configuration space of a set of two particles following a curve
defined by
{ ( x, y, z ) R 3 : f ( x, y, z ) = 0, g ( x, y, z ) = 0 }?

A priori two particles of R 3 have 3 N = 6 degrees of freedom. But the constraint equations
are m = 4 in number because the coordinates of particles ( x1 , y1 , z1 ) and ( x 2 , y 2 , z 2 ) must
verify the following equations:
f ( x1 , y1 , z1 ) = 0 ,

f ( x2 , y2 , z 2 ) = 0 ,

g ( x1 , y1 , z1 ) = 0 ,

g ( x2 , y 2 , z 2 ) = 0 .

The configuration space has thus 3 N 4 = 2 dimensions, the system of two particles has two
degrees of freedom.
More quickly, the curvilinear coordinates of every particle allow seeing that the configuration
space is R 2 such that q1 and q 2 are the curvilinear coordinates of the respective particles.
Example 12. A bike moving along a road has seven degrees of freedom. A priori, the frame
has six degrees of freedom, the back wheel has one, the handlebar has one and the rotation of
the front wheel leads to one in addition. But the wheels must be in contact with the road and
thus there are two constraints, namely: the distance between each wheel center and the
corresponding contact point on the road remains constant.

Statics

75

Example 13. Given two fixed points o and o , a thin rigid rod ab can move in space so that
the distances oa and ob remain constant.
Find the number of degrees of freedom for the rod ab if oabo is such that:
(i)
it remains in a fixed plane,
(ii)
it remains coplanar,
(iii) it does not remain coplanar.

For (i), there is one degree of freedom since each of points a and b has two degrees of
freedom; but there are three constraint equations expressing that the distances oa , ab and
ob are constant.
For (ii), there are two degrees of freedom since the plane can rotate about the axis oo (angle
of rotation!) and thus there is a supplementary degree of freedom with respect to (i).
For (iii), there are three degrees of freedom since each of points a and b has three degrees of
freedom; but there are three constraint equations.

2.1.3

Types of Constraints

We can say:
D

A system of N particles is subject to a constraint if the coordinates of positions of


particles or their first derivatives or both of them are not independent.

Let r1 ,..., rN be position vectors of particles of a system.


D

A constraint is bilateral if the positions or velocities of particles are related by an


equation;1
it is unilateral if the relation is an inequation.

The previous examples show bilateral constraints.


A particle that must stay inside a sphere of radius R is an example of unilateral constraint such
that the inequation is of type
r R < 0,

where r is the radial distance of the particle.


Another example of unilateral constraint is given by two particles a and b, of respective
Cartesian coordinates ( x a , y a , z a ) and ( xb , yb , z b ) , which are connected by an unstretchable
string of length l.
The inequation of constraint is
( xb x a ) 2 + ( y b y a ) 2 + ( z b z a ) 2 l 2 .

For instance :

f (r1 ,..., rN , r1 ,..., rN ; t ) = 0 .

76

Chapter 1

D  A constraint is scleronomic if it is not explicitly dependent on time:


f
= 0.
t

(1-18)

In the opposite case, the constraint is said to be rheonomic.


All the above examples involve scleronomic constraints.
If a pearl follows a ring rotating about a diameter with an angular velocity , then the
constraint is rheonomic since this explicitly depends on the angle t of rotation of the ring.
The simple pendulum with a mass suspended from a moving point gives another example of
rheonomic constraint.
In the same manner, if a pendulum is such that the mass suspended from a fixed point by a
string of not constant length, then the constraint is rheonomic. A crane illustrates this case.
D  A constraint is holonomic if it is expressed by equations of type

f (r1 ,..., rN ; t ) = 0

(1-19)

relating only the coordinates and possibly t.


In the opposite case, it is said to be nonholonomic.
Example 14. A cylinder of radius R is rolling down an inclined plane (without slipping).

If s represents the distance covered by the cylinder and the corresponding angle of rotation
about the cylinder axis, then the constraint is expressed by a relation between velocities
s = R ,

and the constraint is holonomic since this equation is immediately integrated:


s R = k

( k R ).

Example 15. A sphere S of radius R is rolling without slipping on a fixed plane .

Let xc , y c , z c be (primitive) overabundant coordinates of the center c of the sphere, in a frame


of reference
= { o;1x ,1 y ,1z } such that the axes of coordinates x and y are located in .

Let { c;1X ,1Y ,1Z } be a frame fixed in the sphere.


Let , , be Euler angles that specify the position of the frame { c;1X ,1Y ,1Z } with respect
to R .1
The constraint expressing the contact between the sphere and the plane is such that
zc R = 0 .
This is a scleronomic and holonomic constraint.
1

The Euler angles are introduced in first courses of mechanics and are clearly shown in Fig. 43 of Sec. 2.2.3.

77

Statics

Since the sphere is rolling without slipping on the plane (sufficient friction!), there are other
equations of constraint. We express that the slipping velocity at the contact point i is zero:
v i = M i (VS )
= v c + ic
1x

1y

1z

= x c 1x + y c 1 y + z c 1z + 0

R = 0.

x y z
So, we obtain:

x c R y = 0 ,
y c + R x = 0 ,
z c = 0 .
The last equation is the above mentioned holonomic constraint.
We can express the two first equations from Euler angles.
Let 1N be the unit vector along the line of nodes.
We know that
=  1z +  1N +  1Z
where the angular velocities  ,  and  are respectively called the precession velocity, the
nutation velocity and the spin rotation (see also 2.2.3b).
We have:

v i = v c + R 1z ( 1z +  1N +  1Z ) = 0
v + R 1 1 + R 1 1 = 0
c

v c + R 1u + R sin 1N = 0

where 1u is obviously in the plane { i ; 1x ,1 y } and perpendicular to 1N .


Projections onto the x-axis and y-axis lead to the respective equations
x c R ( sin  sin cos ) = 0 ,
y + R ( cos +  sin sin ) = 0 ,
c

which are not integrable. So they represent two nonholonomic constraints.


The six overabundant coordinates verify three equations of constraint and the sphere has thus
three degrees of freedom.
In the preceding, we have assumed that the friction is sufficient to prevent any slipping. We
note that for a frictionless contact the sphere has obviously five degrees of freedom.
Remark. We only consider constraints without dissipation of mechanical energy, every
motion being reversible. The study of constraints which imply a dissipation of mechanical
energy (heat!) is important but the resultant irreversible motions are not here introduced.

78

Chapter 1

2.2

VIRTUAL DISPLACEMENTS AND VIRTUAL VELOCITIES

We are going to introduce the notion of virtual displacement. Such an a priori arbitrary
vector will be particularized so as not to take unknown forces of constraint into account.
Let E3 be a Euclidean vector space,
E 3 be the associated point space (or affine space),
be an open of E3 ,
= { o; e1 , e 2 , e3 } be an orthonormal frame of reference.

2.2.1

Generalized Coordinates

(i)
For a free particle p in E 3 , it is often useful to consider coordinates different from
Cartesian coordinates; so it is better to use three coordinates suited to the geometrical
configuration of the problem (spherical, cylindrical,).
Let U be an open of R 3 ,
q = (q 1 , q 2 , q 3 ) be an element of U composed of generalized coordinates of p.

The position vector of p in R is defined by:


x = r (q (t )) = r (q 1 (t ), q 2 (t ), q 3 (t )) .

(ii)

We consider a particle p belonging to a surface.

Let U be an open of R 2 ,
q = (q 1 , q 2 ) be an element of U, pair of generalized coordinates of p.

A vector function of class C 2 :


r : U : q a r (q ) = r ( q 1 , q 2 )

is a parametric representation of the surface of constraint.


This parametric representation is supposed regular, that is:
(q 1 , q 2 ) U :

where

r
q

and

r
q 2

r
q

r
q 2

are linearly independent vectors and tangent to the surface; the first is

tangent to a curve such that q 2 is constant and the second is tangent to a curve such that q1 is
constant.
Thus there is a tangent plane at each point of the surface.
The position vector of p in R is defined by:
x = r (q (t )) = r (q 1 (t ), q 2 (t )) .

Statics

(iii)

79

We consider a particle p belonging to a curve.

Let U be an open of R,
q be an element of U, generalized coordinate of p.
A vector function of class C 2
r : U : q a r (q)

is a parametric representation of the curve of constraint.


This parametric representation is supposed regular, that is there exists, at each point of the
r
curve, a tangent vector

q
The position vector of p in R is defined by:
x = r (q (t )) .

We consider a free rigid body B in E 3 .

(iv)

Its position is located by six generalized coordinates q j , for instance the three coordinates
x , y , z of the center of mass and the three Euler angles , , .
The position of any point p of B in R is determined by the mapping (supposed of class C 2 ):
r : U ( R 6 ) E3 : ( x , y , z , , , ) a r ( x , y , z , , , ) .
(v)
We consider a mechanical system S made up of k1 rigid bodies, of k 2 rectilinear rigid
bodies and of k 3 particles.
If all these elements are supposed free, then the system has
n = 6k1 + 5k 2 + 3k 3
degrees of freedom.
The position of any point of S in R is defined by the mapping
r : U ( R n ) E3 : (q1 ,..., q n ) a r (q1 ,..., q n )
supposed of class C 2 .
Remark. For every mechanical system in a moving frame of reference and for any system
with rheonomic constraints too, it is necessary to introduce the variable t in addition to the
generalized coordinates.

2.2.2

Definition and Expressions of Virtual Displacements

The following definition concerns general mechanical systems, free or constrained.


Let q 1 ,..., q n be generalized coordinates of a mechanical system S,
U be an open of R n ,
I be an open of R.

80

Chapter 1

We denote
q = (q 1 ,..., q n ) U .

Each particle is located by its position vector x defined by a function of class C 2 :


r : I U : (t , q ) a r (t , q ) .

The components of the position vector x of a mass point p in R are


x1 = r 1 (t , q1 ,..., q n ) ,
x 2 = r 2 (t , q1 ,..., q n ) ,
x 3 = r 3 (t , q1 ,..., q n ) .
Notation. According to usage, the vector function r will be identified with x.

The virtual work method consists in arbitrarily associating with p a neighboring point p .
How? By introducing the notion of virtual displacement, that is an arbitrary and infinitesimal
vector pp resulting from arbitrary increments given to the only generalized coordinates. This
is going to be made more explicit.
We denote this vector by

(or r ).

Arbitrary increments q j of generalized coordinates lead to arbitrary increments of the


components of x, namely:
n

xi =

j =1

r i
qj
j
q

i = 1,2,3.

So, the expression of a virtual displacement of p is


3

r i
q j ) ei
j
j =1 q
n

x = x i e i = (
i =1

i =1

j =1

i =1

= (

r
e ) q j
j i
q
i

and thus


x=

j =1

r
qj.
j
q

(1-20)

We set:
D  A virtual displacement 1 of a point p is a vector which associates with p a neighboring
point from the variation of the only generalized coordinates.

The concept of virtual displacement was introduced by Jean Bernoulli in 1717.

Statics

81

A virtual displacement is a comparison between two neighboring positions at a same instant.


It is a vector that not necessarily represents a displacement!
A virtual displacement is performed at a fixed time. It is a fictitious displacement that would
have been perhaps preferentially called Bernoulli-Lagrange vector although the vector
notion was introduced later by Hamilton.
A fixed time is imposed at any virtual displacement. This requirement will be ever-present in
mind.
We emphasize that a virtual displacement of p is denoted by x and not by dx, this last
being a (real) infinitesimal displacement. Unlike virtual displacements, (real) displacements
take place during a time interval for which the given applied forces and constraint forces
change.
We note that , which acts on generalized coordinates, behaves as a first order differential
operator.
In a general way, the notion of virtual displacement has nothing to do with the one of (real)
displacement. Moreover, there is an infinite number of virtual displacements of p. However, it
is not impossible that a virtual displacement is a real displacement. But, in statics, we notice
that the only possible displacements are virtual displacements!
In the case of systems with constraints, the virtual displacement must be compatible with the
constraint(s) (also called consistent with the constraint(s)). This notion will be specified in the
following.
For example, these virtual displacements are perpendicular to external reactive forces (e.g.
tangent to smooth connections), preserve the rigidity of solids, etc. For a pivot, the virtual
displacement of a point of one body is identical to the one of the contact point of the other
body.
Remark. In this chapter, most of the problems are solved by using generalized coordinates;
the use of (overabundant) primitive coordinates would amount to define virtual displacements
by giving arbitrary increments to primitive coordinates. This will later be made.

2.2.3

Virtual Velocity and Examples

According to usage, every mechanical system S (point, rigid body, system of rigid
bodies) is identified with the part of space occupied by the system at a given time.
2.2.3a Free Particle
In the case of a unconstrained motion of a particle p, the consideration of three
primitive coordinates goes with the number of degrees of freedom. These coordinates are
generalized coordinates.
The particle position in a (continuously) moving frame of E 3 depends on time t and
generalized coordinates q j . It is determined by the vector function

R 4 E3 : (t , q1 , q 2 , q 3 ) a r (t , q1 , q 2 , q 3 ) .

82

Chapter 1

Instead of the notion of virtual displacement, we can consider the concept of virtual velocity.
We recall that the expression of the velocity of p is
vp =

r r 1 r 2 r 3
q& + 2 q& + 3 q& .
+
t q 1
q
q

As for virtual displacements, we consider a fixed time.


So, we can define the notion of virtual velocity of p denoted by v p :
D  A virtual velocity of p is an arbitrary vector
3

v p =

j =1

r j
q&
q j

(1-21a)

where the various q& j are arbitrary reals.


Notation. The arbitrary reals q& j will be denoted by q j and thus a virtual velocity of p is
expressed as
3
r j

q .
(1-21b)
v p =
j
q

j =1
Remark.

We specify if the coordinates q j have the length dimensions, then the

corresponding q j have the velocity dimensions.


In the same manner, if the coordinates q j are angles, then the corresponding q j have the
angular velocity dimensions.

2.2.3b Unconstrained Rigid Body

The consideration of six primitive coordinates goes with the number of degrees of
freedom. These coordinates are the generalized coordinates q 1 ,..., q 6 .
The position of a rigid body B in an (inertial) frame of reference
determined by a function of six generalized coordinates

R = { o;1 ,1 ,1 } is
x

R 6 E3 : (q1 ,..., q 6 ) a r (q1 ,..., q 6 ) .

In some problems, we know that the time t explicitly appears in addition.


To introduce the notions of virtual displacement and virtual velocity, we recall that the
position of the rigid body B is located by six generalized coordinates, which are the three
coordinates x, y, z of a point O fixed in the moving rigid body B and the Euler angles , and
which specify the position of the frame {O ;1X ,1Y ,1Z }, fixed in the rigid body, with
respect to the frame of reference R.

The reader will use the following representation of frames.

83

Statics

Fig. 43

The position of any p B is defined by


R 6 E3 : ( x, y, z , , ,) a r ( x, y, z , , , ) .
Every position vector of p is expressed as:
r = oO + Op
= x 1x + y 1 y + z 1z + X 1X + Y 1Y + Z 1Z .
We immediately have:
r
q 1

r
= 1x ,
x

r
q 2

r
= 1y ,
y

r
q 3

r
= 1z .
z

We are going to calculate the other partial derivative vectors.


We have:

1X = cos 1N + sin 1 y
= cos (cos 1x + sin 1y ) + sin (cos 1y + sin 1z )

but
1 y = sin 1x + cos 1 y

thus
X 1X = X (cos cos sin cos sin ) 1x
+ X (cos sin + sin cos cos ) 1 y + X sin sin 1z .

In the same manner:


1Y = sin 1N + cos 1 y

Y 1Y = Y (sin cos + cos cos sin ) 1x


+ Y ( sin sin + cos cos cos ) 1 y + Y cos sin 1z

84

Chapter 1

and also
1Z = cos 1z sin 1 y
Z 1Z = Z sin sin 1x Z sin cos 1 y + Z cos 1z .

Then, we immediately obtain:


r
r
=
= X [( cos sin sin cos cos )1x + (cos cos sin cos sin )1 y ]
q 4
+ Y [(sin sin cos cos cos )1x + ( sin cos cos cos sin )1 y ]
+ Z [(sin cos 1x + sin sin 1 y ]

= 1z Op
and similarly:
r
q

r
q

r
= 1N Op ,

r
= 1Z Op .

In conclusion, in the case of a free rigid body, the six vectors

r
= 1y ,
y

r
= 1x ,
x

r
q j

are determined by:

r
= 1z ,
z

(1-22)

r
= 1z Op ,

r
= 1N Op ,

r
= 1Z Op .

By taking account of the preceding, a virtual displacement is expressed as follows:


6

x=

j =1

r
q j.
j
q

We can consider virtual velocities instead of virtual displacements.


By considering a fixed time, we define:
D

A virtual velocity of p B is an arbitrary vector


6

v p =

j =1

r j
q
q j

where the various q j are arbitrary reals (also denoted q& j ) and r may explicitly
depend on t.

85

Statics

A field of virtual velocities on B is defined by the mapping

B E3 : p a v p =

j =1

r j
q .
q j

Example. We consider a rigid body B which rotates about a fixed point o.


Let

R = { o;1 ,1 ,1 } be a fixed frame of reference,


x

{ o;1X ,1Y ,1Z } be a frame fixed in the moving body.

The position of the frame fixed in the body is determined with respect to the frame of
reference R by the Euler angles.
The position vector of p B being constant in the frame fixed in the solid, the field of
velocities of p is defined by:
p B : v p = op .

Since

= & 1z + & 1N + & 1Z ,

we have:

(1-23)

v p = & 1z op + & 1N op + & 1Z op .

But the followings


r
= 1z op ,

r
= 1N op ,

r
= 1Z op

imply
r & r
r
v p = &
+
+ &

From this expression, we introduce the notion of virtual velocity as follows.


Given arbitrary reals , and , we consider
v p =

r
r
r
+
+

= ( 1z + 1N + 1Z ) op .

By defining
D

The vector of virtual rotation of B is

= 1z + 1N + 1Z ,

(1-24)

the virtual velocity is expressed as


v p = op .

(1-25)

86

Chapter 1

2.2.3c Systems of Free Bodies

We consider a mechanical system S made up of free bodies. If there are k1 solids, k 2


rectilinear solids and k 3 particles, we know that the system has n = 6k1 + 5k 2 + 3k 3 degrees
of freedom.
Let p be a mass point of S.
By considering a fixed time, we define:
D

A virtual velocity of p is an arbitrary vector


n

v p =

j =1

r j
q
q j

where the various q j are arbitrary reals.


A field of virtual velocities on S is defined by the mapping:
n

S E3 : p a v p =

j =1

r j
q .
q j

2.2.3d Particle Subject to a Constraint

For instance, we deal with a particle moving along a circle which rotates about a
vertical diameter.
In more concrete terms, we consider a small ring (particle p) frictionless running along a
circular rail of radius R and center o. The rails rotates about a vertical diameter with a
constant angular velocity .

Fig. 44
Let { o;1x ,1 y ,1z } be a fixed frame of reference such that 1z is collinear to the vertical
diameter.

Statics

87

Let be the (counterclockwise) angle (1z , op) ,


1 be the unit vector tangent to the circle, at p, of positive sense (counterclockwise),
1 be the unit vector perpendicular to the plane of the circle and such that (1z ,1 ,1 ) is
right ordered,
1r be the unit vector in the direction of r = op,
be the positive angle (1x ,1 ) .
We assume that 1x and 1 coincide at t = 0 , that is

= t.
The small ring has one degree of freedom and we can choose the angle giving the position
of p as generalized coordinate.
By definition, we know that the velocity of transport of p is the absolute velocity of p
considered as fixed in the moving frame.
The expression of this vector is obviously:

R& sin 1 .
Hence, the absolute velocity of p is (with & = ):
v p = R& 1 + R sin 1 .
This vector is obviously not tangent to the circular rail.
A (real) infinitesimal displacement of p is
dop = R d 1 + R d sin 1 .

We are going to search for virtual displacements and virtual velocities.


First of all let us introduce primitive coordinates, namely the radial distance r, the colatitude
(or nutation angle) and the longitude (or precession angle). These coordinates are evidently
not independent since there are equations of constraint:

=t .

r=R

In this example, the Eulerian spin angle (or angle of proper motion previously denoted ) is
zero and 1 is directed along the line of nodes.
So, in a general way, if we give arbitrary increments to primitive coordinates, a virtual
velocity of p is written a priori as follows:
v p =

r r r
r +
+
r

and a virtual displacement as follows:

x=

r
r
r
r + +
r

where r , , , r , and are arbitrary reals.

88

Chapter 1

Since the position vector op of p is indiscriminately written:

x = r = r1r ,
from (1-22) we deduce the followings:
r
= 1r ,
r
r
= 1 r 1r = r 1 ,

r
= 1z r 1r = r sin 1 .

A priori and in an general manner we can thus write:

v p = r 1r + r 1 + r sin 1
and

x = r 1r + r 1 + r sin 1
where the respective terms of this sum are called the radial, meridian and longitudinal terms.
In this problem with one degree of freedom for which is the generalized coordinate, any
virtual displacement or virtual velocity consistent with the constraint has the direction of 1 .
These vectors are respectively written:

x = R 1
and

v p = R 1 .
We also say that these virtual displacement and virtual velocity are compatible with the
constraint equations r = R and = t .
Remark 1. The virtual displacements (resp. virtual velocities) consistent with the constraint
are obtained from virtual displacements (resp. virtual velocities) written with the help of
primitive coordinates; that is:

x = r 1r + r 1 + r sin 1 ,
v p = r 1r + r 1 + r sin 1 .
How? Immediately if we derive the two constraint equations r = R and = t , the time
being fixed! This leads to

r = 0,

= 0

r = 0 ,

= 0.

or

So, virtual displacements and virtual velocities consistent with the constraint are really
expressed as
x = R 1 ,
v p = R 1 .

Statics

89

Remark 2. From the previous classical example, we really notice that an infinitesimal
displacement dx (resp. velocity v p ) of a particle p is not necessarily a virtual displacement
(resp. virtual velocity) consistent with the constraint.
This example confirms the fact that the set of virtual displacements (resp. virtual velocities)
does not necessarily contain the real displacements (resp. velocities). Indeed, the (real)
displacement
R 1 + R sin 1

is not of the following type:


R 1 .

2.2.3e Systems of Constrained Bodies

We consider a system S of rigid bodies.


Primitive coordinates u 1 ,..., u N give the position of S. So, the position of some p S is
determined by a vector function
R N +1 E3 : (t ; u 1 ,..., u N )  r (t ; u 1 ,..., u N )
of class C 2 .

If the system S has n degrees of freedom, we recall that the expressions of any virtual
displacement or any virtual velocity are respectively:
n

x=

j =1
n

v p =

j =1

r
qj,
j
q
r j
q
q j

where the various q j are the n generalized coordinates.


The existence of constraints between solids implies that the N primitive coordinates u i are
not independent, they are not generalized coordinates.
(i)

First, we consider a system S subject to a holonomic constraint of equation:


f (t , u 1 ,..., u N ) = 0 ,

the function f being supposed of class C 2 .


So we have
N
f
f
+ i u i = 0 .
t i =1 u

Before introducing virtual displacements or virtual velocities consistent with constraint


equations, we consider again the example of a small ring following a circle which rotates
about a vertical diameter (see Sect. 2.2.3d).

90

Chapter 1

The constraint equations being


g ) t = 0 ,

f ) r R = 0 ,

we have immediately the following implications:


f
r =0
r

r = 0 ,

g
=0

= 0.

This example shows that the expressions of general virtual displacements (and virtual
velocities) are made simpler since constraint equations exist, these vectors being written:

x = R 1
and
v p = R 1 .

In the case of a holonomic constraint defined by


f (t , u 1 ,..., u N ) = 0 ,

we say:
D

Virtual displacements (resp. virtual velocities) on S are consistent or compatible with


a constraint of Eq. f = 0 if reals u i (resp. u i ) verify the following:
N

u i u i = 0

[ resp.

u i u i = 0 ].
i =1

i =1

If one primitive coordinate, for instance u N , is expressed in the following explicit form:
u N = (t , u 1 ,..., u N 1 ) ,

(1-26)

where is of class C 2 , then we express:


D

Virtual displacements (resp. virtual velocities) on S are consistent (or compatible)


with the constraint of Eq. (1-26) if:

uN =

N 1

i =1

ui
i
u

( u N =

N 1

i =1

i
u ).
u i

(1-27)

So, for S, given holonomic constraints, the field of virtual velocities is


N

S E3 : p  v p =
i =1

(ii)

r i
u .
u i

Second, we consider a system S subject to nonholonomic constraints.

We assume that S is subject to the nonholonomic constraint of equation


N

ai (t , u 1 ,..., u N ) u i + b(t , u1 ,..., u N ) = 0


i =1

ai , b R .

(1-28)

91

Statics
D

Virtual displacements (resp. virtual velocities) on S are consistent with a constraint of


Eq. (1-28) if reals u i (resp. u i ) verify the equation
N

ai u i = 0

ai u i = 0 ) .

(1-29)

i =1

i =1

Now, we assume that S is subject to s nonholonomic constraint equations:


N

a ji (t, u1 ,..., u N ) du i + b j (t , u1 ,..., u N ) dt = 0

(1-30)

i =1

with j = 1,..., s ; a ji , b j R .
We say:
D

Virtual displacements (resp. virtual velocities) are consistent with a nonholonomic


constraint of Eqs. (1-30) if reals u i (resp. u i ) verify the s equations:
N

a ji (t , u1 ,..., u N ) u i = 0

a ji u i = 0 )

i =1

(1-31)

i =1

with j = 1,..., s.
In conclusion, by combining the definitions of virtual displacements (resp. velocities) subject
to both holonomic and nonholonomic constraints, the reader will obviously define fields of
virtual displacements (resp. velocities) in the case of systems subject to any constraints.
Remark. We note that the constraints are holonomic if the differential system (1-30) is
completely integrable; that is, if there are s functions j such that:

j (t , u 1 ,..., u N ) = c j
or

d j =

j
t

j = 1,..., s

i =1

dt +

du i = 0

(so written in the form of nonholonomic constraint equations).

2.2.4

Virtual Fields and Dynams


We consider a (free) rigid body B located by six generalized coordinates q j .

2.2.4a Field of Moments of Dynam


PR27 The vector field of virtual displacements (or virtual velocities) of any p B :
n

x=

j =1

r
qj
j
q

is the field of moments of a dynam.

v p

j =1

r j
q )
q j

92

Chapter 1

Proof. First, we prove that the vector field


B E3 : p 

r
q j

is equiprojective.
Indeed, the distance between any two particles of the moving solid B remains constant, that is:

k R+

a, b B : ab = k

q j R :

ab . (

rb
q

q j

ra
q j

ab

=0=

ab
q j

ab

)= 0.

This equation really expresses that the vector field p 

r
q j

is equiprojective and thus it is a

field of moments of a dynam.


Therefore, each following vector field ( j = 1,..., n) :
p

r
q

qj

(resp. p 

r
q

q j )

and thus their linear combinations, as x (resp. v p ), are also fields of moments of a dynam.
A virtual displacement of a rigid body consists in virtually moving each of points of this
solid; that is, there is a virtual displacement for each point. We must particularize these virtual
displacements so as to conserve the rigid body. Later we will say that such virtual
displacements characterize a rigid body motion.
In order to express such particular virtual displacements, we first introduce the notion of
dynam of virtual velocities.

2.2.4b Dynam of Virtual Velocities and Rigid Body Motion

What is the dynam of which the field of moments was just mentioned?
Let O be an arbitrary reference point fixed in a rigid body B,
p be some material point of B.
The fields of velocities of points of a rigid body are continuous1 since we have:
p, p B : v p v p = p p

v p v p p p .

A vector field w, defined at the instant t, is continuous on B if, for any neighboring points p and
have :

R+ : w ( p ) w ( p ) < .

p of B, we

Statics

93

The well-known formula of field velocities:


p B : v p = v O + pO

leads to the following expression of the virtual velocity:


p B : v p = v O + pO .
It is the moment of a dynam whose resultant is the vector which is the velocity of virtual
rotation of B also called the virtual angular velocity vector.
D

The dynam of virtual velocities or virtual kinematic dynam is defined by its following
elements of reduction


.
v p
From the viewpoint of virtual displacements, we say:
PR28 Every virtual displacement of p B is composed of
- a vector of virtual translation denoted by O,
- a vector of virtual rotation corresponding to a virtual angular displacement
expressed, with respect the basis of the frame of reference, as

= 1 e1 + 2 e 2 + 3 e 3 .
We denote such a virtual displacement by

x = O + pO .
We note the consisting of dimensions.
D

The dynam of virtual displacements of every p B is defined by its following


elements of reduction (at p):

x .

We say:
PR29 Every field of virtual displacements (resp. velocities) restricted to B and associating to
p the vector
n
n
r
r j
j

x=

q
(resp.
q )
=
v

p
j
j
j =1 q
j =1 q
defines a dynam.
Finally, we say:
D

A field of virtual displacements (resp. velocities) on a solid B, at t, is a rigid body field


if it is the moment of a dynam.

94

Chapter 1

So, the following proposition is obvious:


PR30 Every field of virtual displacements (resp. velocities) on B:
n

p x =

j =1

r
qj
j
q

( p  v p =

j =1

r j
q )
q j

is a rigid body field on B (at t).


This notion can be extended to systems of points and solids by considering the restriction of
field of virtual displacements (or virtual velocities) to each of subsystems.

2.2

VIRTUAL WORK

2.3.1

Definitions, Rigid Body Motion and Ideal Constraint

First we are going to define the virtual work and the virtual power of concentrated and
distributed forces.
We consider a mechanical system S in E 3 .
2.3.1a Definitions
D  The virtual work (resp. virtual power) done by a force f acting on a particle p for the
field of virtual displacements x (resp. field of virtual velocities v ) is the real



(resp.

W = f . x ,

(1-32)

P = f . v ).

(1-33)

If the virtual work is not done by any concentrated force f during the virtual displacement but
on the contrary by any distributed force, then we consider the following.
Given a distributed force of continuous density , we consider at any mass point p of an
elementary open of S (of volume d ), the force ( p) d as concentrated at p. So, we say:
D

The virtual work (resp. virtual power) done by a distributed force of density on S
for the field of virtual displacements x (resp. field of virtual velocities v ) is

W = . x d ,

(1-34)

P = . v d ).

(1-35)

(resp.

In the case of concentrated forces f h of various p h S , the virtual work is written:

W = f h . x h .
h

95

Statics

In the case of forces distributed over any subsystem of S, with a force density , the virtual
work is written:

W = d . x .

By denoting df = d , we gather the two previous notions together in only one writing:


W = df . x .

(1-36)

In the same manner, the virtual power of concentrated forces, that is:

P = f h . v h
h

and the virtual power of distributed forces are notions that we can put together in the
following writing:


P = df . v .

(1-37)

2.3.1b Rigid Body Fields

We consider a rigid body field of virtual displacements (resp. of virtual velocities).


We know that a dynam, the resultant of which is (resp. ), is automatically associated
Let q be an arbitrary point of reference. The virtual displacement of p S is such that:

p = q + pq
and
v p = v q + pq .

The corresponding virtual power is written:


P = v p . df = v q . df + qp . df
S

v q .

S df + . S qp df .

But, the dynam of forces is defined at point q as follows:

qp df
S
q

[F ] =

S df

and the dynam of virtual velocities is defined, at point q, by



.

v q

[V ] =

96

Chapter 1

Therefore, the expression of P is a product of two dynams; more precisely, given an


arbitrary point q, we have:
PR31 The virtual power of forces acting on the mechanical system S is the product of the
dynam of forces and the dynam of virtual velocities:

[ ]

P = [F ]. V .

(1-38)

The reader will immediately deduce a similar statement about the virtual work since we can
write, given an arbitrary point q:

W = q . df + . qp df
S

W = [F ] . [ x ] .

(1-39)

2.3.1c Ideal Constraints


D  A constraint between rigid bodies is ideal if the virtual work (resp. virtual power) done
by constraint forces for every field of virtual displacements (or virtual velocities)
consistent with the constraint is zero.

So, in the case of rigid bodies and interconnected systems of rigid bodies, the virtual
displacements and virtual velocities are particularized so as to be compatible with the (ideal)
constraints.
For a rigid body B rotating about a fixed point, any virtual displacement compatible with the
constraint is necessarily the null vector at this point.
Any other material point p h of the solid will have a consistent virtual displacement of
rotation about the fixed point.
This is mathematically expressed as:
O B : O = 0

x h = ph O .

We also recall that if a connection constraints two rigid bodies to have a contact point
continuously, then the reactive force of constraint (internal to the system) does not contribute
to the virtual work if a same virtual displacement is chosen at the common point.
PR32 A constraint corresponding to frictionless contact is ideal.
Proof. Let q be a point of contact between two rigid bodies B1 and B2 .
Let be the common tangent plane at q.
Since there is no friction, the contact dynam at q has a resultant perpendicular to .
But every virtual velocity v q compatible with the constraint has, at q, a direction of the

tangent plane and thus the scalar product of the resultant of the dynam and v q is zero.
So, the corresponding virtual power (or work) is zero; the constraint is ideal.
The converse of this proposition is false; that is:

Statics

97

PR33 An ideal constraint can be with friction.


Proof. A constraint without sliding, at q, is ideal since the field of virtual velocities
compatible with the constraint is zero: v q = 0 .
Therefore the corresponding virtual power (or work) is zero and nevertheless there is friction
because there is no sliding.
Example 1. Characterize the dynam of virtual velocities and the dynam of constraint forces
of a pivot from the definition of an ideal constraint between two rigid bodies B1 and B2 .

Let 1z be the unit vector in the direction of the existent axis of rotation.
At any point a of the axis, the dynam of virtual velocities is the following:
1z
R :
.
0

a
The dynam of forces exerted by B1 on B2 , at a:
f (12)
(12)
M
a

is opposite to the one exerted by B2 on B1 .


The condition of ideal constraint existence:

R : P = 1z . M a(12) = 0
implies:
M a(12) . 1z = 0 .
In conclusion, the moment of the dynam of constraint forces is perpendicular to the axis of the
pivot. This is a well-known result.
Example 2. Characterize the dynam of virtual velocities and the dynam of constraint forces
of a ball-and-socket joint from the definition of an ideal constraint between two rigid bodies
B1 and B2 .

Let o be the center of the joint.


The dynam of virtual velocities is


0 o

where is an arbitrary vector.


The condition of ideal vector existence:
P = . M o(12) = 0
implies that the moment, at o, of the dynam of constraint forces is zero. This is a well-known
result.

98

2.3.2

Chapter 1
Principle of Virtual Work (First Expression)

The goal of the present developments is to formulate the conditions of equilibrium in


terms of generalized coordinates; that is, without reference to vector coordinates of any
material point of a mechanical system.
Let t be an instant of a given interval I.
2.3.2a Particle

Let us prove the following principle of virtual work.


PR34 A particle at rest, at time t, and subject to ideal constraints is in equilibrium with
respect to a frame of reference iff, for every virtual displacement compatible with the
constraints, the virtual work of the resultant of only given forces acting on the particle
vanishes (at any instant t I ).
In brief:
Let x be the position vector of a particle p of mass m,
F be the resultant of given forces applied to p.
The principle is written for every x compatible with the ideal constraints as follows:


p in equilibrium iff W = F . x = 0 .

(1-40)

Proof. The necessary condition is direct.


Indeed, let f be the resultant of all the forces acting on p:
f = F + L.

The hypothesis of equilibrium of p, namely f = 0 , implies:


f . x = ( F + L). x = 0 .

But, the virtual displacements being compatible with the ideal constraints (that is L . x = 0),
we have necessarily:
F . x = 0 .

Let us prove the sufficient condition.


By assumption, we have:
F . x = 0

and for every virtual displacement x compatible with the ideal constraints we know that
L . x = 0 .

Consequently we have:
F . x + L . x = f . x = 0 .

This last result implies that p is in equilibrium. Indeed, if the particle p was not in equilibrium
then it should necessarily be accelerated since it would automatically leave a position of rest
(occupied at a given time). Thus, we would have for a consistent virtual displacement x :

99

Statics
a . x 0

where a is the acceleration of p.


The obtained result, namely:
ma . x 0

is absurd since it is opposite to the hypothesis.


Consequently, the particle p is necessarily in equilibrium.
Remark. The principle of virtual work can be widespread to systems of particles as solids or
interconnected rigid bodies.
For that, we recall that the resultant f h of forces on each particle p h of a system in
equilibrium is zero. So, for any virtual displacement (which is not even compatible with the
constraint!) the virtual work of all forces vanishes in the case of equilibrium:

W = f h . x h = 0 .
h

However, in the principle of virtual work, the virtual displacements are particularized in order
to cancel the virtual work of unknown forces! This last particularity confers great power on
the method!
First, we consider the case of rigid body.

2.3.2b Rigid Body


Preliminary proposition
PR35 The virtual work of internal forces done on a rigid body B is zero.
Proof. The virtual work of internal forces f h(i ) done on a rigid body is

W (i ) = f h(i ) . x h = f h(i ) . O + f h(i ) . ph O


h

=
h

f h(i ) . O

f h(i )

p h O .

where O is an arbitrary point fixed in the rigid body.


We have thus obtained:

W (i ) = f (i ) . O + M O(i ) . .
This means that the virtual work of internal forces is the product of dynams of internal forces
and virtual displacements:

(i )

f (i )
=
. .
(i )
M O O

But the dynam of internal forces of every solid is zero, which implies:

W (i ) = 0 .

100

Chapter 1

In the same manner, we have:

P ( i ) = 0 .
In conclusion, the internal forces do not contribute to the virtual work done on a rigid body
for virtual displacements consistent with the rigidity (that is, in accordance with the existence
of forces which hold the particles of the solid at fixed distance from one another). Now, we
can state the following
Principle of virtual work

PR36 A rigid body B at rest, at time t, and subject to ideal constraints is in equilibrium with
respect to a frame of reference iff, for every virtual displacement compatible with the
constraints, the virtual work of the resultant of given forces done on the rigid body
vanishes (at any instant t I ).
In brief:
Let p h be a particle of the solid,
Fh(e) be the resultant of given forces applied to p h .

The principle is written for every x h compatible with the ideal constraints as follows:


B in equilibrium iff W ( e) = Fh(e ) . x h = 0 .

(1-41)

The reader will state this principle in the power context.


Example. Find again the equilibrium conditions for a rigid body B rotating about a fixed
point O (ball-and-socket joint).
Answer. Any virtual displacement of some p h B :

x h = O + ph O
is chosen compatible with the constraint, that is zero at the fixed point O.
Since there is no possible virtual translation, we have necessarily:

x h = ph O .
The principle of virtual work means that

W = Fh(e ) . x h = Fh( e) . ( ph O )
h

Fh(e )

p h O . = 0

and thus

Oph Fh(e) = M O(e) = 0 .


h

101

Statics
2.3.2c System of Rigid Bodies

In the case of systems of rigid bodies, it is necessary to take given internal forces Fh(i )
into account (if existence). For example, such a force is due to a spring of given torsional
stiffness at the hinge.
We know that the virtual displacements are chosen compatible with the ideal constraints. For
example, a constraint which forces two rigid bodies to keep a common contact point imposes,
for each solid, identical virtual displacements at the common point. We can state:
Principle of virtual work

PR37 A system S of rigid bodies at rest, at time t, and subject to ideal constraints, is in
equilibrium with respect to a frame of reference iff, for every virtual displacement
compatible with the constraints, the virtual work of all the given external and given
internal forces done on the system vanishes (at any instant t I ).
In brief:
Let p h be a particle of the system S,
Fh( e) be the resultant of given external forces acting on p h ,
Fh(i ) be the resultant of given internal forces acting on p h .
The principle is written for every x h compatible with the ideal constraints as follows:
S in equilibrium iff W = ( Fh( e) + Fh(i ) ) . x h = 0 .

(1-42)

The reader will express this principle in the power context.


Remark 1. In the case of a particle or a rigid body, the method of virtual work shows no
great advantage with respect to the classical method. But the method of virtual work is
tremendously profitable in the study of interconnected systems of rigid bodies.
Remark 2. If the system of rigid bodies contains also elastic bodies, then we take these ones
into account when we write the internal forces in the expression of the virtual work principle.

Before extending this principle, until now viewed for systems without elastic members, we
recall that if k designates the stiffness of a spring then the spring exerts a restoring force:
F = kx i

where x measures the distance from the relaxed position of the spring (along the direction of a
unit vector i).
The virtual work done by F for a virtual displacement x is expressed as

kx x = (k x 2 2) = Vel
that is the virtual change in potential energy of the spring.
The principle of virtual work done by given external forces and by elastic internal forces
acting on a mechanical system means that, in equilibrium, we have the following relation:

102

Chapter 1

W ( e) + Wel = 0
where Wel represents the virtual work done by elastic members.
This is obviously written:

W ( e) = Vel .

(1-43)

So, we say:
PR38 A system of interconnected rigid bodies constrained by elastic internal forces is in
equilibrium iff the virtual work done by all given external forces during any virtual
displacement compatible with the constraints equals the virtual change in the elastic
energy of the system.
Remark 3. In any study of equilibrium in a reference system moving with respect to an
inertial frame of reference, it is necessary to take account of forces of transport in the virtual
work expression, but not of Coriolis forces [see Remark in Sect. 1.3.1a].

2.3.2d Torricelli Theorem

PR39 In gravitational field, any virtual displacement1 of the center of mass G of a rigid body
in equilibrium is horizontal.
Proof. Given any particle p h of mass mh and a corresponding (compatible) virtual
displacement x h , in equilibrium, the virtual work is

W = mh g . x h = 0 .
h

But the definition of the center of mass which occupies two neighboring positions G and
G leads to
M GG = mh p h p h .
h

Denoting a virtual displacement of G by G , the expression of virtual work becomes:


M G.g = 0

which proves the theorem.


PR40 In gravitational field, the center of mass G of a rigid body, at rest, fixed at a point
belongs to the vertical line through this point necessarily.
Proof. This proposition follows from the previous theorem. Indeed, if G did not belong to the
vertical line then any virtual displacement consistent with the fixed point would not be
horizontal. Such a conclusion would be absurd seeing the previous theorem.
To conclude this section, we emphasize the following remarks.
1

Compatible with the constraints.

Statics

103

Remark 1. We recall that the generalized coordinates q j do not necessarily have the
dimensions of length (L), they may be angles for instance. However, virtual work (scalar
product) has the dimensions of work, that is ML2T 2 and has the units of force times
displacement, namely: newton times meter (called the joule).
The virtual work can be the scalar product of any force and vector of dimension L. However,
we know that a virtual displacement can be a virtual rotation of a body. In this case, the
virtual work done by the moment M of a couple (M being a free vector of type r F ) during
a virtual angular displacement is

W = M .

(1-44)

The virtual work has the dimensions of work; the units of M are the ones of the work (Nm)
and the angular displacement is expressed in radians.
Remark 2. The number of independent virtual displacements equals the number of degrees
of freedom of any mechanical system. But the reader will be careful of the relations of
dependence between virtual displacements if necessary.
Remark 3. The essential advantage of the method of virtual work is that no reference is
made to unknown reactive forces of constraint. But, if we want to know these forces as well
as the equilibrium positions of a particle or a mechanical system, it is necessary to make the
corresponding constraint forces virtually work. In this case, the virtual displacements are not
chosen consistent with the constraints! This is illustrated with the following example which
also compares the classical and virtual work methods.
Example. A homogeneous ladder ab of mass m and length l rests against an inclined wall
(angle as shown in Fig. 45 ) and on a horizontal floor, the surfaces being smooth.
(i) Let us calculate the horizontal force F, along the x-axis and acting on a, which is able to
hold the ladder in equilibrium; first by the classical method and second by the method of
virtual work.
(ii) Let us find again the reactive forces of constraint from the method of virtual work.

Fig. 45
This problem has one degree of freedom. We choose as generalized coordinate the angle
between the ladder and the floor.

104

Chapter 1
l

(i) The weight mg j of the ladder is applied at point G ( cos l sin cot , sin ) .
The force F = F i is applied at point a( l cos l sin cot , 0 ).
The reactive force of constraint at point a is La = La j .
The reactive force of constraint a point b( l sin cot , l sin ) is Lb = Lb sin i + Lb cos j .
The first condition of equilibrium, namely the resultant of forces is zero, leads to the
respective equations of projection:
Lb sin F = 0 ,
Lb cos mg + La = 0 .
In the same manner, the total moment, for instance at a:
M a = aG m g + ab Lb
introduces a third equation, namely:
l

mg cos Lb l cos( ) = 0 .
2

In conclusion, we have:
F=

mg cos
sin ,
2 cos( )

La = mg ( 1
Lb =

cos cos
),
2 cos( )

mg cos

2 cos( )

Second, any virtual displacement of the ladder, compatible with the constraints, implies that
a is directed along ox and b along ob since the constraints are smooth. This implies:
La . ra = 0 ,
Lb . rb = 0 .
The conditions of equilibrium of the ladder follow from the principle of virtual work for
which only the given active forces must be considered:

W = m g . rG + F . ra = 0 .
We have

ra = xa i
and

rG = xG i + yG j
where the first term of rG does not contribute to W .
Since
l

yG = sin
2

105

Statics

and since the component of rG along j is positive whenever increases by , then we


deduce:
l

y G = cos .
2

Since
x a = l cos l sin cot
and since the component of ra is negative whenever increases by , then we deduce:

x a = (l sin + l cos cot ) .


The condition of equilibrium is thus
l
(mg j ). ( cos j ) + ( F i ) .(l sin l cos cot ) i = 0
2

sin sin + cos cos


l
mg cos ) = 0 .
sin
2

( Fl

The equation holds for arbitrary which implies:


F=

mg cos sin

2cos( )

(ii) In order to calculate the reactive forces of constraint from the method of virtual work, we
are going to make the forces of constraint virtually work.
To obtain La , we choose a virtual displacement compatible with the constraint at point b and
which makes the reactive force La virtually work. For example, we choose 1 as virtual
displacement tangent to the circle of center b, this vector corresponding to a virtual rotation of
(positive) angle about b.
For this virtual displacement, the condition of equilibrium is written:
l

W = m g . 1 + La . l 1 + F . l 1
2

= ( mg cos + La l cos F l sin ) = 0


2

and thus
1

La = mg + F tan = mg (1 +
= mg (1

sin sin
)
cos( )

cos cos
).
2 cos( )

Similarly, to obtain Lb , we choose 1 as virtual displacement tangent to the circle of


center a, this vector corresponding to a virtual rotation of (positive) angle about a.
For this virtual displacement, the condition of equilibrium is written:

106

Chapter 1
l

W = m g 1 + Lb . l 1
2
l

= (mg cos Lb l cos( )) = 0


2

and thus
Lb =

mg cos

2cos( )

This example shows that the method of virtual work used in the case of a unique rigid body
unlike the one of interconnected systems of rigid bodies is not very profitable with respect to
the classical method.

2.3.3

Principle of Virtual Work (Second Expression)

2.3.3a Generalized Force and Principle of Virtual Work


(i) Particle
Given n generalized coordinates, we know that any virtual displacement of a particle p is
defined by
n

x=

j =1

r
qj
j
q

and the virtual work of a force F is


n

W = F . x = F
j =1

r
q j.
j
q

D  The generalized force associated with the generalized coordinate q j and relating to F
is the function

R 2 n +1 R : (t , q, q& ) a Q j (t , q, q& )
such that


Qj = F

r
q j

(1-45)

We assume that the virtual displacements are compatible with the constraints.
Therefore, the principle of virtual work is written:
n

W = Qj q j = 0
j =1

Qj = 0

j = 1,..., n.

(1-46)

The previous equivalence follows from the independence of the n generalized coordinates.
The n equations (1-46) determine the n values of generalized coordinates in equilibrium.

107

Statics

So, we can state the second formulation of the principle of virtual work:
PR41 A particle at rest at a given instant and subject to ideal constraints is kept in
equilibrium with respect to an inertial frame of reference iff all the generalized forces
are zero.

(ii) System of particles


Let rh be the position vector of any particle p h of a system of n degrees of freedom.
Any virtual displacement of p h is defined by
n

rh

j =1

xh =

qj

and the virtual work of the forces acting on the system is

W = ( Fh( e) + Fh(i ) ) . x h
h
n

rh

j =1

= ( ( Fh(e ) + Fh(i ) )

) q j .

The generalized force associated with the generalized coordinate q j and relating to
the given forces acting on the system is the function
R 2 n +1 R : (t , q, q& ) a Q j (t , q, q& )
defined by the scalar product:
Q j = ( Fh(e ) + Fh(i ) )

r
q j

(1-47)

We assume that the virtual displacements are compatible with the constraints.
Therefore, the principle of virtual work is written:
n

W = Qj q j = 0
j =1

Qj = 0

j = 1,..., n.

(1-48)

The previous equivalence follows from the independence of the n generalized coordinates.
The n equations (1-48) determine the n values of generalized coordinates in equilibrium.
So, we can state the second formulation of the principle of virtual work:
PR42 A system of particles at rest at a given instant and subject to ideal constraints is kept in
equilibrium with respect to an inertial frame of reference iff all the generalized forces
Q j are zero.

108

Chapter 1

Remark 1. We point out that the forces of transport must be taken into account if the frame
of reference is not inertial.
Remark 2. Since the various generalized coordinates q j have not necessarily the dimensions
of length, then the generalized forces Q j have not necessarily the dimensions of force.

More precisely, if q k has the dimensions of length, then the dimensions of Qk are MLT 2 .

If q k is an angle, then the dimensions of Qk are ML2T 2 .


So, the virtual work has always the dimensions of work.

2.3.3b Conservative Force Fields and Principle of Virtual Work

(i) Particle
Let (e1 , e 2 , e 3 ) be an orthonormal basis of a Euclidean space E3 .
We recall that a given force field F is conservative iff there exists a continuously
differentiable function named the potential, namely1:
V : E 3 R : p a V ( p)
such that
F = V ,

or similarly:
iff there is a force function U = V such that
F = U .

The expression
F=

U
U
U
e + 2 e 2 + 3 e3
1 1
x
x
x

implies:
Qj = F

r
q j

U
U
U
x1
x 2
x 3
= ( 1 e1 + 2 e 2 + 3 e 3 ) ( j e1 + j e 2 + j e 3 )
x
x
x
q
q
q
=

U x1
x1 q j

U x 2
x 2 q j

U x i

i =1

x i q j

=
=

U
q j

U x 3
x 3 q j

In rheonomic problems, potentials depend explicitly on time.

109

Statics

In conclusion, in the conservative case, the generalized forces are expressed as




Qj =

U
q

q j

(1-49)

This result is valid in rheonomic problems [where explicitly: Q j (t , q )] .


So, we state the principle of virtual work:
PR43 In the case of conservative given forces, a particle is kept in equilibrium with respect
to an inertial frame of reference iff the potential is a function V such that


V
q j

j = 1,..., n .

=0

(1-50)

(ii) System of particles


The resultant of given forces acting on any particle p h is written1:
Fh = hV

and thus
Q j = Fh
h

V
q j

rh
q

= hV
h

rh
q j

Therefore, we state the principle of virtual work:


PR44 In the case of conservative given forces, a system of particles is kept in equilibrium
with respect to an inertial frame of reference iff the potential is a function V such that:


V
q j

=0

j = 1,..., n .

(1-51)

Sometimes these equations are said to define the conditions for a stationary value of V.

2.3.3c Types of Equilibrium

(i) Particle

Let f be the resultant of forces acting on a particle,


dr be any slight displacement (or differential displacement) away from the equilibrium
position toward a neighboring point and taking the constraint into account.

Gradient at point p h : hV = (

V
x

V
x

V
x

).

110

Chapter 1

An equilibrium position of a particle is said to be stable if the work done by f acting


on the particle along any differential displacement dr is negative:
dW = f . dr < 0 ;

it is said to be unstable if:


dW = f . dr > 0 ;

it is said to be neutral if:


dW = f . dr = 0 .

In other words, an equilibrium position is stable iff the vector of projection of f along any
differential vector dr is directed in the sense opposite to dr as shown as follows:

Fig. 46

In the same manner, the reader will consider the unstable and neutral cases.

(ii) Conservative mechanical systems

First, we consider the case of systems having one degree of freedom.


Let q be the generalized coordinate.
The condition of equilibrium is
dV
= 0.
dq

This requirement defines the condition for a stationary value of V; that is, a minimum, a
maximum, a stationary point of inflection or a constant value, in the relation of V versus q.
This type of problem being studied in courses of differential calculus, we only consider the
example of a cylinder on a surface which clearly illustrates the different types of equilibrium
positions.

Statics

111

Fig. 47

In the first case, any slight displacement away from the equilibrium position implies an
increase of potential energy ( V > 0 ), so the cylinder will go back to its equilibrium position
where V is minimum. It is a position of stable equilibrium.
In the second case, any slight displacement away from the equilibrium position implies a
decrease in potential ( V < 0 ), so the cylinder will irremediably move away from the
equilibrium position where V is maximum. It is a position of unstable equilibrium.
In the third case, any slight displacement away from the equilibrium position (inflection
point) means the cylinder will definitely leave the equilibrium position. It is a position of
unstable equilibrium.
In the last case, any slight displacement away from the equilibrium position does not lead the
cylinder to move away from the previous equilibrium position or to go back to this. Any new
position is an equilibrium position. This case, where V = 0 , represents a neutral
equilibrium.
The sign of V is obtained from a Taylor expansion.
Let q0 = 0 be the value of the general coordinate at the equilibrium position. We mention that
if this value is different from zero, then a new variable will be chosen, namely: q = q q 0 .
At a neighboring point of the equilibrium position, the potential is expressed as:
V (q ) = V0 + (

dV
d 2V q 2
d 3V q 3
)0 q + ( 2 )0
+ ( 3 )0
+
3!
dq
2!
dq
dq

where the subscript zero refers to the value of each function at q0 = 0 .


In equilibrium, the condition
(

dV
)0 = 0
dq

implies
V = (

d 2V
dq 2

)0

q2

d 3V

q3
+ ( 3 )0
+
2
3!
dq

112

Chapter 1

The smallness of q implies that the sign of the previous expansion is the one of the lowest
order nonzero term which remains. If it is the second term, then:
d 2V

(i)

The equilibrium is stable ( V > 0 ) iff (

(ii)

The equilibrium is unstable ( V < 0 ) iff (

(iii)

dq 2

)0 > 0 .

d 2V
dq 2

)0 < 0 .
d nV

d 2V

) 0 = 0 , then the sign of the nonzero derivative ( n ) 0 of the lowest order


dq
dq 2
must be examined.

If (

- If this order n is even, then we conclude:


d nV
Stable equilibrium iff ( n ) 0 > 0
dq
d nV

)0 < 0 .
dq n
- If this order n is odd, then we conclude:

Unstable equilibrium iff (

Unstable equilibrium (inflection point).


Second, we consider a system possessing two degrees of freedom.
The sign of the increment V is obtained from the Taylor expansion for the function V of
two generalized coordinates q1 and q 2 . Indeed, let (0,0) be the equilibrium position, after
changes of q1 and q 2 if necessary.
In the neighborhood of the equilibrium position, the expansion for V is
V (q 1 , q 2 ) = V0 + (

) q1 + (
1 0

) q2 +
2 0

( A(q1 ) 2 + 2 Bq1q 2 + C (q 2 ) 2 ) +

where
A=(

2V
(q1 )

B=(

) ,
2 0

2V
q 1q

) ,
2 0

C=(

2V
(q 2 ) 2

)0 .

In equilibrium, the conditions


(

V
q

)
1 0

=(

V
q 2

)0 = 0

imply:
1

V = ( A(q 1 ) 2 + 2 Bq1q 2 + C (q 2 ) 2 ) +
2

By neglecting the derivatives of order higher than 2, then the previous equation becomes the
one of a surface, by expressing V as a function of two independent variables q1 and q 2 .

Statics

113

The differential calculus leads to the following conclusions:


- Stable equilibrium ( V > 0 ) iff B 2 AC < 0 and A + C > 0 .
- Unstable equilibrium ( V < 0 ) iff B 2 AC < 0 and A + C < 0 .
- Unstable equilibrium (saddle point) iff B 2 AC > 0 .
- Undetermined equilibrium iff B 2 AC = 0 .

Fig. 48
The equilibrium is indifferent if V is constant ( V = 0 ).
Example. A mass m slides in a vertical cylinder and is connected by a negligible mass rod of
length l to an end of a spring of stiffness k which slides in a horizontal guide. The other end of
the rod is aligned with the cylinder axis when the spring is unstretched.

Fig. 49
We assume there is no friction and we are going to characterize the equilibrium positions.
The system has one degree of freedom, the angle between the rod and the cylinder axis is
the chosen generalized coordinate.
The spring is unstretched when = 0 o .

114

Chapter 1

From a suitable choice of constants, the potential of the system is


k
V = l 2 sin 2 + mg l cos .
2

The condition of equilibrium:


dV
= k l 2 sin cos mg l sin = 0
d

is fulfilled for the values:

= 0o

and

= cos 1 (

mg
kl

).

The type of stability depends on the sign of


d 2V
= k l 2 (cos 2 sin 2 ) mg l cos .
2
d
For = 0 o , the equilibrium is stable if k >
For = cos 1 (

mg
kl

mg
mg
and is unstable if k <

l
l

) , the equilibrium is stable if k <

mg
mg
and is unstable if k >

l
l

3. EXERCISES
Exercise 1.

Is a vertical wheel of radius R, rolling (without slipping) on a horizontal plane oxy,


subject to holonomic constraints?
What is the number of degrees of freedom of this wheel which cannot pivot?

Fig. 50

115

Statics

Answer. We choose four overabundant coordinates, namely:


- the coordinates xc , y c of the wheel center,
- the angle between the (horizontal) axis of the wheel and x-axis,
- the angle of rotation of the wheel about its axis.
Since the wheel never slips, we have the well-known speed of the contact point:
v = s& = R& .

The components of the velocity of the wheel center are:


x& c = R& sin ,

y& c = R& cos

or in an equivalent manner:
dxc R sin d = 0 ,

dy c R cos d = 0 .

The angle varies and thus the previous equations cannot be integrated; in consequence, the
constraints are nonholonomic.
There are two ( 4 2 ) degrees of freedom, the angles and are the generalized coordinates
for instance.
In this case, the two previous equations determine the overabundant coordinates xc and y c
(if the wheel does not slip!)
Exercise 2.

Find the number of degrees of freedom of a rear axle for which the two wheels, of
radius R, can neither pivot nor slip on a plane.
Characterize the constraints by choosing five overabundant primitive coordinates that are the
coordinates x m , y m , z m of the middle m of the axle, the angle of rotation of wheels and the
angle indicating the direction of the axle.
Answer. The contact points of wheels are i and j, their distance is equal to 2l .

Fig. 51

116

Chapter 1

Since the wheels are rolling without slipping, the following velocities vanish:
v i = v m + im = 0 ,
v j = v m + jm = 0

(1)
(2)

where the angular velocity vector is a priori:

= & 1z + & 1ij


and
im = R 1z + l 1ij

jm = R 1z l 1ij

and thus:

im = ( R& l&)(1vm ) ,
jm = ( R& + l&)(1 ) .
vm

The x,y and z-components of (1) and (2) are immediately:


x& m ( R& l&) sin = 0 ,
y& ( R& l&) cos = 0 ,
m

x& m ( R& + l&) sin = 0 ,


y& ( R& + l&) cos = 0 ,
m

z& m = 0 .
The integration of the last equation leads to the (constraint) equation:
zm = R
which means that the wheels stay in contact with the plane oxy.
The other equations:
x& m = ( R& l&) sin = ( R& + l&) sin ,

y& m = ( R& l&) cos = ( R& + l&) cos

imply obviously:

& = 0

that is the equation of constraint:

= c1

( c1 R ).

Therefore, we have the equations:


x& m = R& sin ,

y& m = R& cos

which could have been directly obtained.


Since the angle is constant, we deduce:
x m = R sin + c 2 ,
where c2 , c3 R .

y m = R cos + c3

117

Statics

In conclusion, the five primitive coordinates are linked together by four scleronomic and
holonomic constraint equations which are:
zm = R ,

= c1 ,

x m = R sin + c 2 ,

y m = R cos + c3 .

The system has one degree of freedom.


Exercise 3.
A disk is rolling and pivoting, but without slipping, on a plane in (known) translation.
Characterize the virtual velocities and virtual displacements which are compatible with the
constraints.
Answer. Let be a plane in given translation,
D be a disk of radius R,
i be the point of contact of D with ,
= { o;1x ,1 y ,1z } be a frame of reference,

R D = { c;1X ,1Y ,1Z } be a frame fixed in the disk where c is the center of D and
1Z is perpendicular to the plane of D.
The plane being in translation, we choose the vector 1z perpendicular to .
Let h(t ) be the height of in R.
We recall that the field of velocities in the motion of translation is constant at any time. The
vector of this field, at instant t, is denoted by
w (t ) = w x (t ) 1x + w y (t ) 1y + h&(t ) 1z .

We choose the following primitive coordinates:


- the coordinates x, y and z of the center c,
- the Euler angles , and specifying the position of basis vectors of R D with respect to
the ones of R.
The nutation angle (or angle of inclination) is such that ] 0, [ .

Fig. 52

118

Chapter 1

A first equation of holonomic and rheonomic constraint is obviously:


z = h(t ) + R sin

and it expresses the permanence of contact.


In addition and in general, we recall that the velocity of some point p of D is
v p = v i + pi + w

where w represents the velocity of transport corresponding to the translation of .


But in this example, the disk is rolling without slipping, thus there is no slip velocity; that is,
the velocity v i (relative to ) of the point of D in contact with at instant t, is zero.
So, by making reference to point c, the slip velocity is
v i ) v c + ci w = 0

where v c is the absolute velocity of c (that is relative to R).


The (scalar) projection of terms of the vector equation onto the unit vector 1 along the
tangent to D at point i leads to the following
v c . 1 + (& 1z + & 1 + & 1Z ) ci . 1 w . 1 = 0

x& cos + y& sin + R& cos + R& w x cos w y sin = 0 ,

that is the equation of a nonholonomic constraint.


In the same manner, the (scalar) projection of terms of the already used vector equation onto
the unit vector 1n = 1z 1 leads to the following
v c . 1n + (& 1z + & 1 + & 1Z ) ci . 1n w . 1n = 0

x& sin + y& cos + R& sin + w x sin w y cos = 0 ,

that is the equation of a nonholonomic constraint.


So, the disk D has three ( 6 3 ) degrees of freedom.
The most general fields of virtual velocities and virtual displacements are respectively:
p a v p =

r r r r r r
x +
y + z +
+ +
,
x
y
z

pax =

r
r
r
r
r
r
x + y + z + +
+
.
x
y
z

They are compatible with the constraints for x , y , z , , and verifying the following
relations:
z = R cos
[because (1-27)],
x cos + y sin + R cos + R = 0

[because (1-29)],

x sin + y cos + R sin = 0

[because (1-29)].

Statics

119

Exercise 4.

A hoist is a system of two pulleys having several grooves, the upper pulley being fixed
and the lower being moving. A weight W is acting on the lower pulley. An inextensible rope
passes n times over pulleys, one of its ends is fastened to the lower pulley, while at its other
end a vertical force F is exerted down.
Determine the force F required to hold W in equilibrium.
Answer. This system has one degree of freedom.
By assuming there is no friction, the only forces are the reactive force L exerted on the fixed
pulley, the given force W and the force F.

Fig. 53
Any virtual displacement compatible with the constraints must maintain fixed the upper
pulley in order to eliminate the contribution of L to the virtual work.
To a vertical virtual displacement of length y of the lower pulley corresponds a virtual
displacement of length n y along the line of action of F. Indeed, the length of the rope
passing over the pulleys being constant, then the virtual displacement associated with F is
divided between n parallel segments of the rope.
The chosen virtual displacement of the point of action of F has the same directional sense as
F and the directional sense of the corresponding virtual displacement of W is thus opposite to
the directional sense of W.
Thus, by introducing the magnitudes F and W, the equilibrium condition is written:
W y + n F y = 0 ,

that is:
F = W
n
Exercise 5.

A rope, passing over a vertical pulley on which a mass m is hung, is fixed to a ceiling
at one of its ends. The other end p is fastened to a spring which is fixed to the ceiling. The
radius of the pulley is R and the spring stiffness is k.
If the pulley is released initially from the position where the force in the spring is zero,
calculate the angle of rotation of the pulley when the equilibrium is reached.

120

Chapter 1

Answer. In this problem of one degree of freedom, let y be an arbitrary increase of the
ordinate of the center c of the pulley.
The virtual displacement of c being vertical and down, the virtual work done by the force
acting on the pulley is
m g . rc = mg y .

Fig. 54

Consequently, the fastening point p undergoes a virtual displacement of increase ( 2 y ) since


the point i represents the instantaneous center of zero velocity (Fig. 54).
Then, the virtual work done by the (elastic) restoring force is written:

Wel = k 2 y (2 y ) = 4ky y
but

y = R

and thus the following condition of equilibrium

W = mgR 4kR 2 = 0
implies:

mg

4kR

Exercise 6.

Given a weight W exerted on a truss at a point p as shown in Fig. 55, express the
magnitude F of the force Fq or Fs in the top member of the truss composed of equilateral
triangles, the length of each structural element being l.
Answer. First, we recall that in the classical method we must be careful to the constraints,
namely here a pin connection, at o, which is capable of supporting a force in any direction in
the plane normal to the pin axis, while the rocker at point a can support a vertical force only.
But the method of virtual work allows to find F without such considerations.

The friction forces are neglected.

Statics

121

Fig. 55

The angle is the chosen generalized coordinate for the framework, this problem having one
degree of freedom.
We choose the virtual displacements compatible with the constraints; namely: there are no
virtual displacement of the fixed point o, no deformation of structural elements and no
friction; it is so that we consider any increase .
The condition of equilibrium is written:

W = Fq . rq + Fs . rs + W . r p = 0 .
The only horizontal components of rq and rs must be taken into account.
Since x q = l cos , the (negative) component of rq in the i-direction is x q = l sin
and the (positive) component of rs in the i-direction is x s = l sin .
The one and only component of r p is the (positive) j-component:

y p = l .
In conclusion, the principle of virtual work is written:
(W j ) . (l j ) + ( F i ) . (l sin i ) + ( F i ) . (l sin i ) = 0

W l + 2 Fl sin = 0

that is
F=

2 sin

122

Chapter 1

Exercise 7.
A slider-crank mechanism is composed of a rod op of length R, rotating about a fixed
point o, and a rod pq of length l where p represents the crank pin.1 The particle q slides along
a horizontal rail.
All the frictions are neglected.
What force F along the rail is sufficient:
(i)
to hold a weight W in equilibrium at p;
(ii)
to hold in equilibrium the moment M of a couple about o; more precisely deal with
this question when a piston P is pinned to the connected rod and slides in a cylinder.
(iii) Find again the result of this last case by introducing virtual velocities.

Fig. 56
Answer. (i) The chosen generalized coordinate of the mechanical system of one degree of
freedom is the angle between the rail and the rod op.
Any virtual displacement must take the fixity of o into account; that is, o = 0 . We have thus
for the reactive force L at o:
L . o = 0 .

Since o is fixed and the rod op is rigid, any virtual displacement p compatible with the
constraints is necessarily tangent to the circle of radius op = R .
In addition, the reactive force of constraint acting on the particle q is perpendicular to the rail,
so that any consistent virtual displacement q will be horizontal.
Thus the principle of virtual work leads to

W = W . p + F . q = 0 .
Now, we calculate the only component of q as follows.
To the abscissa of q:

x q = R cos + l cos = R cos + l 1 ( R l ) 2 sin 2


(where is the angle between the connecting rod and the rail)
corresponds the following virtual change:
1

In French, the crank op is called the manivelle and pq is called the bielle.

123

Statics

sin cos

R2
x q = ( R sin
l

1 ( R l ) 2 sin 2

) .

This expression is negative for any increase and thus we have


R2
l

q = ( R sin

sin cos
1 ( R l ) 2 sin 2

) i .

The virtual work done by W is only due to the vertical component of any virtual displacement
p.
Since the ordinate of p is y p = R sin , the j-component of p is

y p = R cos .
The vertical component y p is positive for any increase , so that the principle of virtual
work is immediately written:
sin cos

R2
(W j ) . ( R cos j ) + ( F i ) . ( R sin
l

1 ( R l ) 2 sin 2

) i = 0

W cos + F sin (1 +

R
l

cos
1 ( R l ) 2 sin 2

) = 0 .

The increase being arbitrary, we deduce:


F=

R
sin (1 +
l

W cos
cos

1 ( R l ) sin
2

In particular, if R = l , we have:
F=

W
cot .
2

(ii) A piston P slides in a horizontal cylinder as shown in Fig. 57.


If we consider the force f tangent to the circle (o, R) , then the moment M of the couple is
M = op f = R f k .

From previous comments about reactive forces of constraint, we can state here the principle
of virtual work as follows:
Mk . k + ( F i ) . x q i = 0

M + FR sin (1 +

cos
R
) = 0 .
l 1 ( R l ) 2 sin 2

124

Chapter 1

Fig. 57
Thus the equilibrium exists for
F=

R
cos
R sin (1 +
)
l 1 ( R l ) 2 sin 2

f
R
cos
)
sin (1 +
l 1 ( R l ) 2 sin 2

We point out that:


The previous result is immediately obtained by expressing the virtual work done by f, that is:
f . s = fR

where s is any consistent virtual displacement of p.


The dimensions of the virtual work done by M are really the ones of a work since the chosen
virtual displacement is a virtual angular displacement k .
By introducing virtual velocities, the principle of virtual work takes the following form:

f . vp + F. vq = f vp F vq = 0
and thus we have:
F= f

v p
v q

= f

ip
iq

From
ip
sin(90 )
o

iq
sin( + )

we deduce:
F=

f cos
=
sin( + )

f cos
f
=

sin
R cos
sin (cos +
cos )
sin (1 +
)
sin
l cos

Statics

125

Exercise 8.

The platform of mass m of a vehicle goes up by the application of the moment M of a


couple acting about the lower end a of a jointed rod of length l = 8 (m). There are horizontal
slots which allow the linkage to unfold as the platform is elevated.
In equilibrium, express the norm of M as a function of the height h of the platform. Calculate
this norm M if the platform 4 meters high has a mass of 8 tons. Any friction is neglected.

Fig. 58
Answer. The system has one degree of freedom and the chosen generalized coordinate is the
angle of inclination of the rod at a.
The frame of reference { a; i , j , k }, the weight mg and the moment M are shown in Fig. 58.
The condition of equilibrium is

W = mg . rG + M k . k = 0
where rG and k are virtual displacements compatible with the constraints.
Since the weight m g is vertical, only the j-component of rG matters, namely:
l sin + k

where k is the height of G with respect to the platform.


For any (positive) increase , the virtual work done by m g is negative and the one of M is
positive. We have thus:

W = mg l cos + M = 0
which implies

M = mg l cos

= mg l 1 (h l ) 2
= 543171 (Nm).

126

Chapter 1

Exercise 9.

An industrial tool fastened to a plate goes up and down when a handle rotates. A
rotation of constant angular velocity (uniform rotation) generates a uniform translation of the
plate through mechanisms not to be described. When the handle, the arm of which is 0.18 (m)
in length, turns once then the plate goes up of 0.1 (m).
What force F perpendicular to the handle is necessary to hold the plate and tool of 250 (N)
weight in equilibrium?
Answer. This system has one degree of freedom. We choose a virtual displacement
compatible with the constraints such that to any increase y of the height of the plate
corresponds an increase of angle . These increases are linked by the relation:

y
=

2
0 .1
The principle of virtual work, which is written as follows:

W = 0.18 F 250 y = 0

(0.18 F

2
0 .1

250) y = 0 ,

leads to the result:


F = 22.1 (N).

This exercise proves again the efficiency and the simplicity of the method of virtual work
since it is not necessary to describe the (unknown) internal mechanisms.

Exercise 10.

Two bars oa and ob of length l, in a vertical plane, are pivoted about a pin at fixed
point o. An interconnected system of four vertical bars ad, db, bc, ca of same length L is
jointed to the first two bars at a and b. All the massless bars, in the same plane oxy, are
connected by frictionless pins.
Given a force Fd applied at d, find the force Fc to apply at c in order to hold the system in
equilibrium.
Answer. A set of four moving particles in a plane has eight degrees of freedom a priori; but
here the number of degrees of freedom is 2 because there are six constraint equations.
Let { o;1X ,1Y } be the moving frame such that the X-axis is through c and d.
We choose the following generalized coordinates:
- the angle which determines the direction of oX relative to the x-axis,
- the angle between oa (or ob) and the X-axis.
We are going to introduce virtual displacements compatible with the constraints.
On the one hand, we choose an arbitrary increase , while remains constant.

127

Statics

Fig. 59
In this case, where the distance between o and c is necessarily constant (and also between o
and d), from
rc = oc 1X ,
rd = od 1X
we deduce respective virtual displacements:

rc =

rc
= oc 1Y ,

rd =

rd
= od 1Y .

Thus the corresponding generalized force is


Q = Fc

rc
r
+ Fd d

that is:
Q = ( Fc ) Y oc + ( Fd ) Y od
where the notation ( ) Y represents the projection onto the Y-axis.
In equilibrium, the condition Q = 0 implies:
( Fc ) Y =

od
( Fd ) Y .
oc

(1)

On the other hand, we choose an arbitrary increase , while remains constant.


In this case, where c and d move closer (along X-axis), we have:

rc =

rc
.

128

Chapter 1

From
rc = ( l cos L2 l 2 sin 2 ) 1 X

we deduce the virtual displacement:

rc = (l sin +

l 2 sin cos
L l sin
2

l sin

L2 l 2 sin 2

) 1X

( l cos L2 l 2 sin 2 ) 1X

and so:
l oc sin

rc =

L2 l 2 sin 2

1X .

In the same manner, since


od = l cos + L2 l 2 sin 2 ,

we have:
l od sin

rd =

L2 l 2 sin 2

1X

and the generalized force is expressed as


Q = Fc

that is
Q =

rc
r
+ Fd d

l sin
L2 l 2 sin 2

( oc ( Fc ) X od ( Fd ) X ) .

In equilibrium, the condition Q = 0 implies:


( Fc ) X =

od
( Fd ) X .
oc

(2)

So we have:
Fd = [( Fd ) X cos + ( Fd ) Y sin ] 1x + [( Fd ) X sin + ( Fd ) Y cos ] 1 y ,

Fc =

od
od
[( Fd ) X cos ( Fd ) Y sin ] 1x
[( Fd ) X sin + ( Fd ) Y cos ] 1 y
oc
oc

and, for instance, Fc is obtained from Fd in equilibrium.

Statics

129

Exercise 11.

A block of mass M supported by two rollers of same mass m is rolling (without


slipping or pivoting) down an inclined plane, being the angle of inclination. We assume the
rollers neither slip nor pivot.
Calculate the force F, applied to the block parallel to the inclined plane, sufficient to hold the
system in equilibrium.

Fig. 60
Answer. The problem would be complex by using the classical method since there are four
reactive forces of friction which prevent slipping.
In this problem of one degree of freedom, the chosen generalized coordinate is the coordinate
x of the center of mass G of the block such that x is positive upwards.
Let us consider virtual displacements compatible with the constraints. At time t, the point i1
of a roller in contact with the incline at fixed point i2 has no velocity since the roller is not
slipping. It is the instantaneous center of zero velocity.
So, the virtual velocities and displacements compatible with the frictional constraints are zero:

ri1 = ri2 = 0 .
It is an example of ideal constraint with friction.
In the same manner, the point j 2 of the block in contact with the point j1 of the roller is such
that:
r j1 = r j2 .
The reactive forces of constraint do no virtual work, and the principle of virtual work is
expressed as follows:

W = Mg . rG + F . rG + m g . rc1 + m g . rc2 = 0
where c1 and c2 belong to respective roller axes.

130

Chapter 1

But for this pure rolling motion, we have at time t:


1

rc1 = r j2 = rG
which is rc2 too.
The virtual displacements are not independent since there is only one degree of freedom. The
principle of virtual work is written:

W = Mg sin xG + F xG mg sin xG = 0
which implies:

F = ( M + m) g sin .

Exercise 12.

Two particles of respective masses m1 and m2 are located on a frictionless double


incline. The constant angle between the inclined planes is denoted by . Both planes rotate
about a smooth hinge, the axis of this pivot being horizontal. The particles are connected by
an inextensible string (of length l and negligible mass) passing over the smooth hinge.
Find the conditions of equilibrium of the particles by using the principle of virtual work.

Fig. 61
Answer. We choose the system of coordinates oxz in the vertical plane of motion.
The system made up of particles has two degrees of freedom. Let us introduce the following
generalized coordinates:
-

the angle between a plane and the vertical as shown in Fig. 61,
the length l1 = op1 .

We have obviously op 2 = l l1 .
The principle of virtual work is written:

W = m1 g . r1 + m2 g . r2
= m1 g z1 m2 g z 2 = 0

131

Statics

with
z1 = l1 cos ,

z 2 = (l l1 ) cos( ) .

The constraints imply the following relations:

(l l1 ) = l1 ,
( ) =
and

z1 = l1 cos + l1 sin ,
z 2 = l1 cos( ) + (l l1 ) sin( )( ) .
The principle of virtual work is written:

W = [m1 g cos m2 g cos( )] l1 + [m1 g l1 sin + m2 g (l l1 ) sin( )] = 0.


Since the general coordinates l1 and are independent, we deduce the two conditions of
equilibrium:
m1 g cos = m2 g cos( ) ,
m1 g l1 sin = m2 g (l l1 ) sin( )
which express that the forces along the string have the same norm and the moments of forces
about the origin o have the same norm.

Exercise 13.

A particle p of mass m, which is travelling along a frictionless vertical circle of radius


R, is the end of a spring; the other end is fastened to the fixed point a ( R,0) in the frame of
reference { o; i , j } in the plane of the circle.
Knowing that the spring of stiffness k is 2 R in length when relaxed, do equilibrium positions
of p exist?

Fig. 62
Answer. The chosen generalized coordinate is the angle = (i , op) . The spring is always
compressed and thus the elongation x is always negative except if ap = 2 R , that is x = 0 .

132

Chapter 1

The restoring force F is directed from point a toward p and is expressed as


F = k ( ap 2 R ) 1ap .
An arbitrary virtual displacement of p will be chosen so that the reactive force of constraint,
which is normal to the circle, does not virtually work. Therefore, such a displacement must be
tangent to the circle.
From

= ap = ( R cos R) i + R sin j

we deduce:

= R sin i + R cos j .

This last vector being perpendicular to the position vector op, it is really tangent to the circle

and so is p =
.

Let us use the second formulation of the principle of virtual work.


From

( mg j + 2kR 1ap k ap)

kR 2 sin
= mgR cos + 2
kR 2 sin = 0
1 cos

Q =

and since
sin = 2 cos

2 sin

= 2 cos

1 cos

we deduce:
Q =

mg

cos sin + 2 cos = 0 .


kR
2

By letting x = cos , the previous equation is written:


2

mg
(2 x 2 1) + 2 x 1 x 2 2 x = 0.
kR

By letting C =

mg
, the following equation
kR
(1 + C 2 ) x 4 2C x 3 C 2 x 2 + C x +

C
=0
4

allows to obtain the equilibrium positions of the particle p.


If we let
f ( ) = C cos + sin 2 cos

we have:

f ( ) = C < 0

Statics

and
f (3 2) = 1 + 2 > 0 ,

which means there is really an equilibrium position for ( , 3 2) .


In the same manner, since
f ( 0) = C 2

and
f ( 2) = 1 2 < 0 ,

if C > 2 , that is if k <

mg
, then there is an equilibrium position for (0 , 2) .
2R

133

CHAPTER 2

TENSORS

The idea of tensor took form at the end of the 19th century, when it was necessary to
express pressure forces in continua. But, the first important developments of the tensor notion
date back in the very beginning of the 20th century; they are generally due to Ricci, LeviCivita, E. Cartan,
The name tensor, introduced by the physicist Voigt, is reminiscent of tension in fluids,
elastic solids, This terminology designates intrinsic mathematical beings, which are suitable
for the expression of laws of mechanics regardless of the choice of coordinate system.
The reader has already encountered tensors since vectors and linear forms are
examples of them.
It is pointless insisting on the considerable importance that the tensors have gained
through developments of exact and applied sciences in the 20th century, more especially in
Riemannian geometry, special relativity, general relativity, quantum mechanics, differential
geometry, analytical mechanics, fluid dynamics, cosmology, electricity, electromagnetism
and so on.1

1. FIRST STEPS WITH TENSORS


1.1

MULTILINEAR FORMS
Let E and F be finite-dimensional real vector spaces.

1.1.1

Linear Mapping

A mapping

g : E F : x  g( x)

is linear if x , y E, k R :
g ( x + y) = g( x ) + g ( y) ,
1

g (kx ) = k g ( x )

This chapter is based on our books: Mcanique Analytique vol.2 (1982) and Differential geometry with
Applications to Mechanics and Physics (2000), this last one introducing tensors in the manifold context.

135

136

Chapter 2

Let us denote by L( E ; F ) the set of (continuous) linear mappings of E to F.


D

The addition in L( E ; F ) is the mapping


L( E ; F ) L( E ; F ) L( E ; F ) : ( g , h)  g + h

such that the sum g + h is the linear mapping defined by


E F : x  ( g + h)( x ) = g ( x ) + h( x ) .

The multiplication of a linear mapping g of E into F by a scalar k is the mapping


R L( E; F ) L( E; F ) : (k , g )  k g

such that the product k g (of g by k) is the linear mapping defined by


E F : (k , g )( x ) = k g ( x ) .

We know that L( E ; F ) provided with the two previous laws of addition and multiplication
has a structure of vector space.
1.1.1 Multilinear Form
In mechanics, we particularize F by choosing this vector space to be R. So, we will
consider the vector space L( E ; R) later on.
D

A linear form on E is a mapping


f : E R : x  f ( x)

such that, x , y E, k R :
f ( x + y) = f ( x ) + f ( y)

f (kx ) = k f ( x ) .

A linear form on E is also called a one-form or covector.


Let E (1) ,, E ( p ) be p vector spaces.
D

A p-linear form defined on the Cartesian product of p spaces E (1) ... E ( p ) is a


mapping
f : E (1) ... E ( p ) R : ( x (1) ,..., x ( p ) )  f ( x (1) ,..., x ( p ) )
which is linear with respect to each vector, that is:
x (1) , y (1) E (1) ,, x ( p ) , y ( p ) E ( p ) , k R :
f ( x (1) + y (1) , x ( 2) ,..., x ( p ) ) = f ( x (1) , x ( 2) ,..., x ( p ) ) + f ( y (1) , x ( 2) ,..., x ( p ) )

...
f ( x (1) ,..., kx ( p ) ) = k f ( x (1) ,..., x ( p ) ) .
Provided with laws of addition and multiplication by a scalar defined as before, the space
L p ( E ; R) of p-linear forms on E has a structure of vector space.

137

Tensors
1.2

DUAL SPACE, VECTORS AND COVECTORS

1.2.1

Dual Space

The vector space of linear forms defined on E is called the dual space of E.

It is denoted by E .
So, the dual space is a vector space the elements of which, called covectors, are linear
functions E R . It is a space of functions.
Example 1. The row vectors are covectors (or 1-forms). With the multiplication of matrices,
a row vector (linearly) associates a real to each column vector.
For instance:
x
( 3 ,1) : R 2 R :  3 x + y .
y
Example 2.

In quantum mechanics, the 1-forms called bras and denoted , linearly

associate complexes , to vectors called kets and denoted .

1.2.2 Expression of a Covector

Let E be a real vector space of dimension n.


A covector on E is a linear mapping f : E R which associates a real f ( x ) to each
vector xE.
We denote by x1 ,..., x n the components of x with respect to a basis (e1 ,..., e n ) of E.
The real f ( x ) is written:
f ( x ) = f ( x1e1 + ... + x n e n )
= x1 f (e1 ) + ... + x n f (e n ).

By letting
f i = f (e i ) ,

we have
n

f ( x) = f i x i .
i =1

We mention that the image of x under f is sometimes called the value of the form (for x).
Now, we are going to express the covector f with respect to the dual basis.
The dual basis (e 1 ,..., e n ) of the basis (e1 ,..., e n ) is such that:


e i (e j ) = ij

(2-1)

138

Chapter 2

where ij is the Kronecker delta1 and the n linear forms making up the dual basis are
e i : E R : x  e i ( x ) = x i .

So, x E :
n

f ( x ) = f i e i ( x ) ,
i =1

which leads to the expression of the covector




f = f i e i .

(2-2)

i =1

Remark. The reader will compare the previous expression with the one of a vector:
n

x = x i ei .
i =1

So, according to usage, the components of vectors show an upper index and the components
of covectors a lower index.
Notation. Generally, we will represent the covectors (or linear forms) by Greek characters
and, since they are the elements of a vector space, namely E , we have decided to write them
in bold characters.

1.2.3

Einstein Summation Convention

The Einstein summation convention consists in removing the summation sign , more
precisely:
Summation is implied when an index is repeated on upper and lower levels.2
For example, we denote:
n

ai x i = ai x i ,
i =1
n

f i e i = f i e i ,
i =1

aijk x i y j z k = aijk x i y j z k .
i =1 j =1 k =1

The Kronecker delta is the symbol

1 if i = j
0 if i j .

ij = ij = ij =
2

The Einstein convention will be also used with indices at the same height.

139

Tensors

On the one hand, any repeated index of summation is called dummy index because no matter
what the letter is like; for instance:
ai x i = a j x j = a1 x1 + a 2 x 2 + ... ,
bij x i x j = brs x r x s = b11 x1 x1 + b12 x1 x 2 + ... ,
aii = a jj = a11 + a 22 + ...

In particular, the Kronecker symbol is such that

ii = 11 + 22 + ... = 1 + 1 + ...
Attention! With the summation convention, we emphasize that an index of summation shall
never be repeated more than once.
So, there is no question of writing:
aij b j x j

(never!),

bii c j x i y j

(never!).

In the same manner, given


Q = aij x i y j
with
y j = bi j x i ,
if we want to express Q in function of coordinates x i only, then there is no question of
writing:
Q = aij bi j x i x i
(never!),
but we must firstly change the dummy index in y j , for instance:

y j = bkj x k
in order to obtain the following right expression:
Q = aij bkj x i x k .
In the same manner, if we want to multiply ai x i and bi y i , we must first rename the dummy
index of one term of the product and write ai b j x i y j .
On the other hand, there is another type of index. An index, which appears once in each
expression, is called a free index.
So, for instance, the equation

y i = b ij x j
represents the following system of equations:
y 1 = bk1 x k ,
y 2 = br2 x r

that is explicitly:

i = 1,2 ; j = 1,2,3,4

140

Chapter 2
y1 = b11 x1 + b21 x 2 + b31 x 3 + b41 x 4 ,
y 2 = b12 x1 + b22 x 2 + b32 x 3 + b42 x 4 .

In this example, the index i is free and the index j is dummy.


Remark1. The reader will take care to write the equations correctly. Any introduced free
index must appear in every term.
So, the following equations are

Meaningless:

meaningfull:

xi = y j

(never!)

zj = xj + yj,

aij = aik

(never!)

a qp = b ip ciq .

Remark 2. Factoring is possible with the Kronecker symbol.


For instance, by viewing ni = i j n j where i j is the Kronecker symbol, the equation

ij n j = ni

is written:
( ij i j ) n j = 0 .

1.2.4

Change of Basis and Cobasis

Let (e1 , e 2 ) be a basis of a 2-dimensional vector space called unprimed basis and
(e1 , e 2 ) be another basis called primed basis to make simpler.

The following change of basis


e1 = 11e1 + 12 e 2 ,

e1 = 11e1 + 12 e 2 ,

e 2 = 12 e1 + 12 e 2

e 2 = 21 e1 + 22 e 2

is written in a condensed manner:


e j = ij e i

( )

ei = ik e k

( )

where = ij and = ik are inverse matrices and thus qp qr = pr .


The equalities
x = x1e1 + x 2 e 2 = x 1e1 + x 2 e 2
= x 1 (11e1 + 12 e 2 ) + x 2 ( 12 e1 + 22 e 2 )

imply by comparison:
x1 = 11 x 1 + 12 x 2 ,
x 2 = 12 x 1 + 22 x 2 .

141

Tensors

In the same manner, the equalities


x = x 1e1 + x 2 e 2 = x1e1 + x 2 e 2
= x1 ( 11e1 + 12 e 2 ) + x 2 ( 21e1 + 22 e 2 )

imply by comparison
x 1 = 11 x1 + 21 x 2 ,
x 2 = 12 x1 + 22 x 2 .

We generalize and, in an n-dimensional space E, we say:


PR1

The matrix associated with the expression of unprimed components in function of


primed is the transpose of the matrix associated with the expression of primed basis
vectors in function of unprimed.

( )

( )

Proof. The matrices ij and ij being inverse, we denote




e j = ij e i ,

(2-3)

ei = ik e k .

(2-4)

The equations
x = x i e i = x k e k = x k ki e i

imply
x i = ki x k .

(2-5)

By comparing the following explicit expressions


x1 = 11 x 1 + 12 x 2 + ... ,

e1 = 11e1 + 12 e 2 + ...

and so on, the proposition is thus proved.


PR2

The matrix associated with the expression of primed components in function of


unprimed is the transpose of the matrix associated with the expression of unprimed
basis vectors in function of primed.
It is inverse and transposed of the matrix associated with the expression of primed
basis vectors in function of unprimed.

Proof. Since
ei = i j e j ,

the equalities
x = x j e j = x i e i = x i i j e j

imply


x j = i j x i .

By comparing the following explicit expressions


x 1 = 11 x1 + 21 x 2 + ... ,
e1 = 11e1 + 12 e 2 + ...

(2-6)

142

Chapter 2
e1 = 11e1 + 12 e 2 + ...

and so on, the proposition is thus proved.


Now, we are going to show the formulae of change of dual bases (a dual basis is also called a
cobasis).
From every
x = x i e i = x j e j ,

because e j (e k ) = kj , we deduce:
e j ( x ) = x j = i j x i = i j e i ( x )
which implies


e j = i j e i

(2-7)

e i = ij e j .

(2-8)

and


The reader will easily say the propositions which refer to (2-7) and (2-8).
Example. Primed axes being obtained from a 90 direct rotation1 about the e 2 -axis, express
the vector 4e 3 with respect to the primed basis ( ei ).

The primed basis vectors e j = ij e i are explicitly


e1 0 0 1 e1


e 2 = 0 1 0 e 2
e 1 0 0 e
3
3
and the unprimed vectors ei = i j e j are explicitly:
e1 0 0 1 e1


e 2 = 0 1 0 e 2 .
e 1 0 0 e
3
3
From PR2, we deduce:
x 1 0 0 1 x1



x 2 = 0 1 0 x 2 .


x 3 1 0 0 x 3

The given vector (such that x 3 = 4 ) is written 4e1 since x 1 = 4 , x 2 = x 3 = 0 .


This result is obvious because e3 = e1 .

A direct rotation is also called a counterclockwise rotation.

143

Tensors
1.3

TENSORS AND TENSOR PRODUCT

Let E (1) , , E ( p ) , , E ( p + q ) (or simply E) be finite-dimensional vector spaces.


1.3.1 Tensor Product of Multilinear Forms

Let f be a p-linear form defined by


E (1) ... E( p ) R : ( x (1) ,..., x ( p ) )  f ( x (1) ,..., x ( p ) ) ,

let g be a q-linear form defined by


E ( p +1) ... E( p + q ) R : ( x ( p +1) ,..., x ( p + q ) )  g ( x ( p +1) ,..., x ( p + q ) ) .

D  The tensor product of a p-linear form f and a q-linear form g is the (p+q)-linear form
denoted f g :
E (1) ... E( p + q ) R : ( x (1) ,..., x ( p + q ) )  f g ( x (1) ,..., x ( p + q ) )

such that
f g ( x (1) ,..., x ( p + q ) ) = f ( x (1) ,..., x ( p ) ) g ( x ( p +1) ,..., x ( p + q ) ) .

(2-9)

Example 1. If we consider two linear forms f and g defined on E, namely:


f : E R : x  f ( x)
g : E R : y  g ( y ),

then the tensor product of linear forms f and g is the bilinear form
f g : E E R : ( x, y)  f g ( x, y ) = f ( x ) g ( y) .

Given a basis (e i ) of E, the linear forms f and g have the respective values:
f ( x ) = f i x i = f i e i ( x ) ,
g ( y) = g j y j = g j e j ( y)
and the corresponding value of the tensor product is
f g ( x, y) = f ( x ) g ( y ) = f i g j x i y j .
Example 2. The tensor product of two bilinear forms

f : E(1) E( 2) R : ( x (1) , x ( 2) )  f ( x (1) , x ( 2) ) ,


g : E ( 3) E ( 4 ) R : ( x ( 3) , x ( 4 ) )  g ( x ( 3) , x ( 4 ) )
is the following quadrilinear form
f g : E(1) ... E( 4) R : ( x (1) ,..., x ( 4) )  f g ( x (1) ,..., x ( 4) )
= f ( x (1) , x ( 2) ) g ( x (3) , x ( 4) ).

144

Chapter 2

1.3.2

Tensor of Type

()

A tensor of type

( ) or covector is a linear form defined on E.

0
1

0
1

It is an element of the vector space E .


According to usage, the covectors are generally denoted by Greek letters; for instance:

E .
The definition of a covector or tensor of type

( ) is thus expressed as follows:


0
1

: E R : x  ( x )
with a, b R, x , y E :
(ax + by ) = a( x ) + b( y ) .
The covector expressed as (2-2) is written:

= i e i
where

i = (e i ).
The image of any vector x under is the real

( x ) = i e i ( x j e j ) = i x j e i (e j )
= i xi ,
this value being also denoted by

, x = ( x ) .

(2-10)

Change of basis

We recall that a linear form behaves towards any vector x in the following way:

( x ) = ( x i e i ) = ( x j e j )

i x i = j x j = , x

(2-11)

where j = (e j ) .
This obvious requirement allows testing the tensor character. Let us use it in order to obtain
the formulae of transformation of components of .
By recalling (2-5) and (2-6):
x i = ki x k ,

x p = np x n ,

y j = rj y r ,

y s = ms y m ,

145

Tensors

the condition (2-11) that is

i x i = i ki x k
= k x k

implies

k = ki i

(2-12)

i = ir r .

(2-13)

and conversely

1.3.3

Tensor of Type

()

A tensor of type

( ) or vector is a linear form defined on E

1
0

1
0

So, the definition of a tensor of type

( ) or vector x is expressed as follows:


1
0

x : E R :  x ()

such that a, b R, , E :
x (a + b ) = ax () + bx ( ) .

A linear form x defining a tensor of type


of E as follows:
x = x i ei
where
x i = x (e i ) .

( ) is obviously written with respect to a basis ( e )


1
0

So, the image of any covector under the linear form x is the real
x () = x i e i ( j e j ) = x i j i j = x i i
which is written:

x () = x, .
We have


x, = , x

and this important result expresses the duality between covectors and vectors.
Remark 1. In fact, we may identify E = L( E ; R) with E:
E = E .

(2-14)

146

Chapter 2

By referring to the formulae of basis change (2-3) and (2-4) in E on the one hand, and to their
corresponding (2-7) and (2-8) in E on the other hand, we immediately see that the
concerned vectors e i of E and ei of E are transformed according to the same rule.
To each vector expressed with respect to a basis ( ei ) of E corresponds a vector of same
components with respect to the corresponding basis ( e i ) of E and conversely, such that to
the sum of any two vectors of E corresponds the sum of two corresponding vectors of E , to
the product of a vector of E by a scalar corresponds the product of the corresponding element
of E by this scalar.
Since there is no reason to distinguish the elements of E from the ones of E, we have the
right to identify these vector spaces.
Algebra courses deal with this question, and the existence of an isomorphism between finitedimensional vector spaces E and E is easily proved.
Remark 2. Following from the duality expressed by (2-1), we point out that the covector e i
of the dual basis associates with x the ith component x i :

e i , x = e i ( x ) = e i ( x j e j ) = x i .
Terminology and notation. We note that the law (2-12) of change of components of any
covector is the one of change of basis vectors (2-3).
It is not the case for a vector: the matrix is inverse! Thats the reason why, initially, every
vector (element of E) was called contravariant vector and every covector (element of E )
was named covariant vector.
This terminology is logically given up because vectors and covectors exist as own entities
regardless of any basis change.
But later, it could well be that we say indices of contravariance and indices of covariance.

According to convention, the components of vectors show an upper index and the components
of covectors a lower index.
Recall that the Einstein summation convention about indices is systematically used and that
any basis vector ei is distinguished by a lower index while any covector e i of the dual basis
is characterized by an upper index.
We are going to consider all that with tensors of higher order.

1.3.4

Tensor of Type

()
0
2

In the next chapter, we will see the inertia tensor as an example of this tensor type.
D

A tensor of type

( ) is a bilinear form defined on E E .


0
2

Sometimes called a covariant tensor of order 2.

147

Tensors
D

The vector space of bilinear forms defined on E E is called tensor product space of
two spaces E .

It is denoted

E E.
So, any tensor of type

( ) is an element of E
0
2

E and we denote such a tensor:

t E E .
Tensor expression. Given a basis ( ei ) of E, we say:
PR3

A tensor of type ( 02 ) is expressed as

t = t ij e i e j

(2-15)

where
t ij = t (e i , e j )

and (e i e j ) is a basis of E E .
Proof. Let x = x i e i be any vector of E where the various x i are defined by1
e i : E R : x  e i ( x ) = x i .

Let us form the n 2 tensor products


e i e j : E E R : ( x , y )  e i e j ( x , y )

such that2
e i e j ( x , y ) = e i ( x ) e j ( y ) = x i y j .

Let us prove that the n 2 tensors e i e j of type

()

( ) compose a basis of E
0
2

E .

First, every tensor t of type 02 is a linear combination of the different elements e i e j .


Indeed, its value for any ( x , y ) is

t ( x, y ) = t ( x i e i , y j e j ) = x i y j t (e i , e j ) = t ij x i y j
= t ij (e i e j )( x, y ).
So, every tensor t of type

( ) is written:
0
2

t = t ij e i e j ,
its components being the reals t ij = t (e i , e j ) .

We also denote

e i , x = x i .

We also denote

e i , x e j , y = x i y j .

148

Chapter 2

Second, the different elements e i e j are linearly independent. Indeed, for every pair of
vectors (e r , e s ) we have:
e i e j (e r , e s ) = e i (e r ) e j (e s ) = ri sj .

Therefore, we have , E :
= i j (e i e j ) = 0

i j (e i e j )(e r , e s ) = 0
r s = 0
r , s {1,..., n}.

Remark. As a vector x is sometimes and excessively called by its components x i , a tensor t


of type 02 can be called by its components t ij (and so for higher order tensors).

()

Change of basis

Remember that any tensor is an intrinsic mathemathical being, that is independent


of the choice of basis; in other words, each real defined by a bilinear form t is not altered by
a change of basis.
Given a change of basis defined by e j = ij e i , the components t ij of a tensor t E E are
transformed as follows:
( x , y ), ( x , y ) E E :

t ( x , y ) = t ( x , y )

t ( x i e i , y j e j ) = t ( x r e r , y s e s )

t ij x i y j = t ' rs x r y s .

(2-16)

Such a (general) requirement of tensor theory allows testing the tensor character. So, let us
use it in order to obtain the formulae of transformation of components of t.
We recall [see (2-5) and (2-6)]:
x i = ki x k ,

x p = np x n ,

y j = rj y r ,

y s = ms y m .

The condition (2-16), namely:


t ij x i y j = t ij ki rj x k y r
xk yr
= t kr

implies
= ki rj t ij .
t kr

(2-17)

Remark - Rule. We will notice the presence of elements of two matrices in (2-17) and of
one matrix in (2-12); that is, to each covariance index corresponds one matrix . Its the
reason why every tensor of type 02 is sometimes called second order covariant tensor.

()

Conversely, we have:
.
t ij = ir sj t rs

(2-18)

149

Tensors
1.3.5

Tensor of Type

()

A tensor of type

( ) is a bilinear form defined on E

The vector space of bilinear forms defined on E E is called the tensor product
space of two spaces E.

2
0

2
0

This space being denoted E E , any tensor of type

E .

( ) is such that
2
0

t E E.

Tensor expression. The reader can fit the previous developments about tensors of type

in the case of tensors of type

( ).

()
0
2

2
0

He will define n linear forms on E :


e i : E R :  e i () = i

and then will consider the n 2 tensor products


e i e j : E E R : (, )  e i e j (, )

such that
ei e j (, ) = e i () e j ( ) = i j .

Therefore, he will be able to state:


PR4

A tensor of type

( ) is expressed as
2
0

t = t ij e i e j

(2-19)

where
t ij = t (e i , e j )

and (e i e j ) is a basis of E E .
For that, the reader will prove that if (e i ) and (e j ) are bases of E, then the different ei e j
are linearly independent.
He will also prove that every tensor t E E is expressed as (2-19).
Change of basis

It is proved (see Exercise 4) that the components of a tensor t of type


transformed as follows:
t rs = ir sj t ij .

()
2
0

are

(2-20)

150

Chapter 2

Remark Rule. We will notice the presence of elements of two matrices in (2-20) and of
one matrix in (2-6); that is, to each contravariance index corresponds one matrix (inverse
of the matrix of basis change). Its the reason why every tensor of type 02 is sometimes
called second order contravariant tensor.

()

Conversely, we have:
t ij = ip qj t pq .

(2-21)

Remark. Given two vector spaces E q and E r of respective dimensions q and r, the
corresponding tensor product space is E q E r of dimension qr. It is the set of tensor
products x y of any x E q and any y E r .

Bases ( ei ) and ( e j ) of respective spaces E q and E r imply that ( ei e j ) is a basis of the


qr-dimensional space E q E r .

1.3.6

Tensor of Type

()

A tensor of type

( ) is a bilinear form

Thus a tensor of type

1
1

1
1

defined either on E E or on E E .

( ) is either an element of the tensor product space


1
1

E E or an

element of the tensor product space E E .


From the covectors of the dual basis:

e i : E R : x  e i , x = x i

and the vectors of the basis of E:


e j : E R :  e j , = j ,

the reader will define the n2 tensor products:


e i e j : E E R : ( x,)  e i e j ( x ,)

such that
e i e j ( x,) = x i j

(2-22)

that is
e i , x e j , = x j e i , e j i e j , e i = x i j .

As before, the reader will establish that the various tensors e i e j form a basis of the vector
space E E (likewise for E E ).
1

Sometimes called mixed tensor of order 2.

151

Tensors

From
t ( ,x ) = t (i e i ,x j e j ) = i x j t (e i ,e j )
= i x j t ij = t ij e i e j ( ,x ) ,
he will deduce
t = t ij e i e j E E
and also
u = u i j e i e j E E .
In addition, he will easily establish the formulae of transformation of the previous tensors:
t r s = ir sj t i j
u r s = ri sj u i j ,
the recalled rule being: To every covariance index corresponds one matrix and to every
contravariance index one matrix .
Given t = t ij e i e j E E , we say:
D

The transposed tensor of t is the tensor of E E , denoted t t , such that


e i E , e j E :
t

t (e i , e j ) = t ( e j , e i ) .

(2-23a)

In other words:
(t t )i j = t j i .

(2-23b)

For instance, given = i e i and x = x j e j , the expression


x = i x j e i e j

leads to
t

( x ) = x i j e i e j = x .

In the same manner, the reader will define the transposed tensor of u E E .

()
q
p

1.3.7

Tensor of Type

The vector space of p-linear forms defined on E ... E (p spaces E) is called the
tensor product space of p identical vector spaces E .

It is denoted
p E
and has np dimensions.

152

Chapter 2

In the same manner, we say:


D

The vector space of q-linear forms defined on E E (q spaces E ) is called the


tensor product space of q identical vector spaces E.

It is denoted
q E
and has n q dimensions.
D

The

( )-tensor space associated with E is the vector space of (p+q)-linear forms


q
p

defined on the Cartesian product ( p E ) ( q E ) of p spaces E and q spaces E .


This n p +q dimensional space is denoted 1

T pq = ( p E ) ( q E )
D

A tensor of type

( ) associated with E is an element

q
p

of the

( )-tensor space T
q
p

q
p .

We denote this (p+q)-linear form by


t T pq .
The reader will easily verify the following proposition.
PR5 A tensor of type

( ) is expressed as
q
p

t = t i1...i p

j1... jq

e i1 ... e

i p

e j1 ... e jq

(2-24)

where
t i1...i p

j1... jq

= t (e i1 ,..., e i p , e j1 ,..., e

and where the different e i1 ... e

i p

jq

e j1 ... e jq constitute a basis of T pq

= ( p E ) ( q E ) .
Change of basis

Every transformation of components of a tensor of type


change (2-3) is obvious by considering the rule:

( ) following from a basis


q
p

[ covariance index

matrix ]

[ contravariance index

matrix ] .

and

To simplify the presentation, we have first chosen p spaces E and next q spaces
spaces must be specified.

Also called a p-order covariant and q-order contravariant mixed tensor.

E . The order of successive

Tensors

153

From this rule, it is easy to express any transformation of components.


For example, let
t ij

e i e j e k e l

k
l

()

be a tensor of type 13 .
To express the primed components in function of the unprimed, we simply proceed as
follows.
Having written
r
k
t pq s = p. q. .r s. t ij l
where the various and follow from the rule, we immediately replace the dots by
successive indices of t ij k l , that is:
t pq

= ip qj kr sl t ij

k
l

Remark 1. According to usage, it is necessary and useful to consider tensors of type


They are the scalars (independent of basis choice!).

( ).
0
0

Remark 2. We recall that the previously introduced Kronecker symbol is only a symbol (and
not a tensor).

But we can introduce:


D

The Kronecker tensor is a tensor of type

ij =

1
0

if
if

( ) of which the components are


1
1

i = j,
i j.

It is a very helpful tensor because its components are unaltered under any basis change:

i j = ir sj rs = ir rj = i j .
D

The zero tensor is a zero multilinear form denoted by 0.

1.3.8

Symmetric and Antisymmetric Tensors

Let us consider for instance elements of E E [resp. E E ].


D

A tensor of type
x , y E :

()
0
2

[resp. of type

( ) ] is symmetric if
2
0

t ( x , y ) = t ( y, x )

[resp. , E :

t (, ) = t ( ,) ].

This definition is equivalent to the following

154

Chapter 2

A tensor t ij e i e j [resp. t ij e i e j ] is symmetric if


t ij = t ji

[resp. t ij = t ji ].

t ( x, y ) = t ( y, x )

t ( x i e i , y j e j ) = t ( y i ei , x j e j )

t ij x i y j = t ij x j y i

t ij = t ji .

Indeed, we have:

Of course, the property of symmetry is intrinsic; that is, independent on the basis: Given a
change of basis e j = ij e i , we have:
t ij = ip qj t pq = ip qj t qp = qj ip t qp = t ji ,
[likewise: t ij = t ji ].
The previous definition may be generalized to higher order tensors.

()

A tensor of type qp is partially symmetric if it is symmetric with respect to pair(s) of


corresponding indices;
in other words:
If there are symmetries following from every transposition of two indices of same
variance.

A tensor of type 0p or type 0q is completely symmetric if every transposition of


indices changes the corresponding component into itself.

()

()

Remark. Given an n-dimensional vector space, every symmetric tensor of order 2 has
n(n + 1) 2 independent components.

Now we consider tensors which play an important role in mathematics and physics: the
antisymmetric tensors.
D

A tensor t of type
x , y E :

( ) [resp. of type ( ) ] is antisymmetric if


0
2

2
0

t ( x, y ) = t ( y, x )

[ resp. , E : t (, ) = t ( ,) ].
This definition is equivalent to the following
D

A tensor t ij e i e j [resp. t ij e i e j ] is antisymmetric if


t ij = t ji

[resp. t ij = t ji ].

155

Tensors

Indeed, we have:

t (, ) = t ( ,)

t ( i e i , j e j ) = t ( i e i , j e j )

t ij i j = t ij j i

t ij = t ji .

Obviously, the property of antisymmetry is intrinsic; that is, independent on the basis: Given a
change of basis e j = ij e i , we have:
t ij = ip qj t pq = ip qj t qp = qj ip t qp = t ji ,
[likewise t ij = t ji ].
We deduce:
t ( x, x ) = 0

or

t ii = t ii = 0 .
The previous definition may be generalized to higher order tensors.

()

A tensor of type qp is partially antisymmetric if it is antisymmetric with respect to


pair(s) of corresponding indices;
in other words:
If there are antisymmetries following from every transposition of two indices of same
variance.

A tensor of type 0p or of type 0q is completely antisymmetric if every transposition


of indices changes the corresponding component into its opposite.

()

()

Remark. Given an n-dimensional vector space, every antisymmetric tensor of order 2 has
n(n 1) 2 independent components.

PR6

()

()

Every tensor of type 02 [or of type 02 ] can always be decomposed, in a unique


manner, into the sum of a symmetric tensor t S and an antisymmetric tensor t A :
t = tS + t A .

Proof. We have, x , y E :
t S ( x, y ) = 12 (t ( x, y ) + t ( y, x )) ,
t A ( x, y ) = 12 (t ( x, y ) t ( y, x ))
and the proof follows.

156

Chapter 2

Given any basis ( ei ) of E and its dual basis ( e i ), the previous proposition is also proved
from
(t S ) ij = 12 (t ij + t ji ) ,
(t A ) ij = 12 (t ij t ji ),

[likewise with tensors of type

( ) ].
2
0

2. OPERATIONS ON TENSORS
Before defining operations on tensors, we say:
D

Tensors are equal if they are the same element of a same tensor space.

Example. The following tensors

t = t i jk e i e j e k ,
u = u p qr e p e q e r
of the same space E E E are equal if all their corresponding components are equal
t=u

t i jk = u i jk

for every index value.

2.1

TENSOR ALGEBRA

2.1.1

Addition of Tensors

An inner law, namely the addition, can be defined on the set of same type tensors.
We say:
D

The sum of two tensors of which the n p + q components are respectively t j11 ... qj p and
i ...i

i ...i

u j11 ... qj p is the tensor of type ( qp ) the components of which are


i ...i

i ...i

t j11... qj p + u j11... qj p .
The addition of two tensors of type ( qp ) is

T pq T pq T pq : (t , u)  t + u
where t + u is the tensor sum.

157

Tensors
2.1.2 Multiplication of a Tensor by a Scalar
D

i ...i

The product of a tensor with components t j11 ... qj p by a scalar k is the tensor whose
components are
i ...i

k t j11 ... qj p .

The multiplication of a tensor of type ( qp ) by a scalar k is

R Tx qp Tx qp : (k , t )  k t
where k t is the product of t by k.
Example. Given two tensors t, u T21 , we have k1 , k 2 R :
k1t + k 2 u = k1t i jk e i e j e k + k 2 u i jk e i e j e k
= (k1t i jk + k 2 u i jk ) e i e j e k .

2.1.3 Tensor Multiplication


D

The tensor multiplication of any tensor t of type ( qp ) and any tensor u of type ( rs ) is the
mapping
: T pq Trs T pq++rs : (t , u)  t u ,
the tensor product t u being such that
t u ((1) ,...,( q ) ,( q +1) ,...,( q + s ) , x (1) ,..., x ( p ) , x ( p +1) ,..., x ( p + r ) )
= t ((1) ,...,( q ) , x (1) ,..., x ( p ) ) u(( q +1) ,...,( q + s ) , x ( p +1) ,..., x ( p + r ) ). (2-25)

Example. t T21 , u T11 :

t u = (t i jk e i e j e k ) (u p q e p e q )
= t i jk u p q ei e j e k e p e q .
It is a tensor of T32 .
This law verifies the following properties:
P1. The tensor multiplication is bilinear:
t T pq , u(1) , u( 2) Trs : t (u(1) + u( 2) ) = t u(1) + t u( 2) ,
t (1) , t ( 2) T pq , u Trs : (t (1) + t ( 2) ) u = t (1) u + t ( 2) u ,
k R, t T pq , u Trs : k (t u) = kt u = t ku .

These properties are immediately checked from the definition of tensor product.

158

Chapter 2

P2. The tensor multiplication is associative:


t , u, s T pq : (t u) s = t (u s ) = t u s.
P3. The tensor multiplication is not commutative.

Let us give the following obvious counter-example of tensor product of vectors:


x y = yx

x i y j ei e j = y j x i e j ei

xi y j = yi x j

xi

xj

= (k )
kR
yi y j
that is iff the two vectors are parallel. So the tensor multiplication is generally not
commutative.

D

The tensor algebra is the infinite-dimensional vector space:


T = R E E T20 T02 T11 ... T pq ... ,

direct sum of vector spaces the dimensions of which are higher and higher, and where
R represents the tensors of type ( 00 ) (also called scalars).
This space is provided with a bilinear inner law: the tensor multiplication.
Therefore, we can express:
PR7

The tensor algebra T is associative, non-commutative and of infinite dimension.

2.2

CONTRACTION AND TENSOR CRITERIA

We consider tensors of type ( qp ) such that p, q 1


2.2.1 Contraction
D

The contraction of a tensor is the operation which consists in choosing a


contravariance index and a covariance index, in equalling these and in summing with
respect to the repeated index.

For example, let us consider the tensor


k

t mp k e m e p e T12 .

Contracting in p and k, we obtain a tensor of T01 whose components are


u m = t mp p = pk t mp k = t m11 + t m 2 2 + ...
p

It is a tensor having lost a contravariance and a covariance because the component change is
such that:

Tensors

159

u m = pk t mp k = pk im jp kr t ij r = im kj kr t ij r = im rj t ij r
= im u i .
This is in accordance with the rule of change of vector components, and thus the contraction
of t T12 leads to u T01 .
We can express:
PR8

Every contraction of a tensor removes one contravariance and one covariance.

PR9

After q contractions a tensor of type ( qq ) is reduced to a tensor of type ( 00 ) (in


principle q! in number).

Example 1. Given a 3-dimensional space E, the contraction of

t = t i j ei e j
leads to the scalar

t i i = t 11 + t 2 2 + t 3 3 .
Example 2. Given a 2-dimensional space E, two successive contractions of
t = t ij e i e j e k e r
kr

lead to the following scalars:

t kr kr = t1111 + t1212 + t 2121 + t 22 22 ,


t rk kr = t1111 + t12 21 + t 2112 + t 22 22 .
D

The contracted multiplication is the tensor multiplication with contraction.

Example. The contracted multiplication

E E R : (, x )  , x
is such that
, x = i x i .

(2-26)

Among the different contractions of a tensor product of two tensors t u , we emphasize the
following:
Notation. The contraction with respect to the last index of t and the first index of u is denoted
by 1
t u.

If this conventional writing is allowed.

160

Chapter 2

Example. Given two tensors


t = t ij e i e j e k
k

u = u qr e q e r ,

the contraction of
t ij u qr e i e j e k e q e r
k

with respect to indices k and q leads to


t u = t ij u kr e i e j e r .
k

Among the possible double contractions of a tensor product of two tensors t u , we


emphasize the following:
The contraction with respect to the last index of t and the first index of u, followed by the one
with respect to the penultimate index of t and the second index of u is denoted by
t : u.

Remark. The dot between two tensors corresponds to the previous type of contraction. If the
contraction concerns other indices, then this must be specified by letting both the indices of
contraction between brackets.

For instance, given t = t ij e i e j and x = x k e k , we have the following covector


t x = t ik x k e i
= (t11 x1 + t12 x 2 + ...)e 1 + (t 21 x1 + t 22 x 2 + ...)e 2 + ...

which is different from the (1,1) contraction:

t kj x k e j = (t11 x1 + t 21 x 2 + ...) e 1 + (t12 x1 + t 22 x 2 + ...) e 2 + ...


Following from the properties of the tensor multiplication, we immediately have:
- the associative property:
(t (1) t ( 2) ) t (3) = t (1) (t ( 2) t (3) ) = t (1) t ( 2) t (3)

and
(t (1) : t ( 2) ) : t 3 = t (1) : (t ( 2) : t (3) ) = t (1) : t ( 2) : t (3) .

- the following distributive property:


(t (1) + t ( 2) ) t (3) = t (1) t (3) + t ( 2) t (3) .

Example 1. Given x = x i e i E , y = y j e j E and t = t pq e p e q E E , we


immediately have:
x t y = t ij x i y j
= t ( x, y)

Tensors

161

since

t pq e p e q ( x, y ) = t ij x i y j .
Example 2. Given x = x i e i E and t = t p e p e q E E , we have:
q

x t = x i ti q e q E

and also
t

t x = (t p q e p e q ) x i e i
= x i t pi e p .

So,
x t = tt x .

(2-27)

Example 3. The double contraction of any tensors t E E and u E E is a scalar


since, given
t = t ij e i e j ,
u = u pq e p e q ,
we have:
t : u = t ij u ji .
It is commutative:
t :u = u:t .
Example 4. Given arbitrary second order tensors t, u and v, verify that
t : ( u v ) = u : (v t ) = v : ( t u) .

For instance, given tensors of type

(2-28)

( ), we have
1
1

t : (u v ) = t qp (u v ) qp = t qp (u ip v qi ) = u ip v qi t qp = u ip (v t ) ip
= u : (v t )

and so on.

Remark 1. The double contraction of tensors of order 2 decomposed into symmetric and
antisymmetric parts is such that:

t : u = t S : uS + t A : u A .
Indeed, we have t E E , u E E :

t S : u A = (t S ) ij (u A ) ji = (t S ) ji (u A ) ij = t S : u A = 0
and

t A : u S = (t A ) ij (u S ) ji = (t A ) ji (u S ) ij = t A : u S = 0 .

162

Chapter 2

Remark 2. The following terminology is used in mechanics of continua.

A scalar associated with a tensor is said to be an invariant if its value is the same in all
coordinate systems.
In second volume in hand, we study invariants of the (second order) stress tensor.
For instance, the trace of a second order tensor t is an invariant denoted by
tr (t ) = t ii

(summation over i).

t pp = ip pj t ij = ip jp t ij = ij t ij = t ii .

(summation over p).

Indeed, we have:

The determinant is another example of invariant which appears in the characteristic equation
of the stress tensor .

2.2.2 Tensor Criterion

Until now, we have been able to recognize tensors either from the definition directly,
or from the transformation of components through basis changes. We are going to show a
very useful criterion ensuring the tensor character of given mathematical entities; it will be
based on contractions.

(i) Special case. We know that given two vectors x = x i e i and y = y j e j , if aij x i y j is a
tensor of type

( ) (scalar), then the various a


0
0

ij

are the components of a tensor.

Conversely, if the various aij are the components of a tensor a, then the expression aij x i y j
which is the result of two contractions, namely between the tensor a and the tensor x y of
type

( ) , is a scalar.
2
0

(ii) Special case. With respect to a basis (e i e j e k ) , we consider n 3 expressions aij k


of which one may wonder if they are tensor components.
First, given arbitrary following vectors and covector x = x i e i , y = y j e j , = k e k , if the
different aij k are the components of a tensor, then aij k x i y j k is an (intrinsic) scalar.
Conversely, if the previous aij k x i y j k is an (intrinsic) scalar, then we are going to prove that
the different aij k are the components of a tensor.
Indeed, let x p , y q and r be primed components of respective arbitrary vectors and
covector with respect to a basis (e j ) = ( ij e i ) .
The scalar is such that
r
k
k
a pq x p y q r = aij x i y j k = a ij ip qj kr x p y q r
and thus, given the arbitrariness of the vectors and of the covector, we have:
r
k
a pq = aij ip qj kr .
This proves that the various aij k are the components of a tensor.

163

Tensors

The previous example may be generalized; n p + q expressions are the components of a tensor
of type qp iff the complete contraction with respect to p arbitrary vectors and q arbitrary
covectors is an (intrinsic) scalar.

()

(iii) General tensor criterion.


First, in particular, we consider n 3 expressions aij k and an arbitrary covector = p e p . We
know that if the various aij k are the components of a tensor, then the various aij k k (= Aij )
are the components of a tensor.
Conversely, if the various Aij are the components of a tensor; then, given arbitrary vectors x
and y, Aij x i y j is an (intrinsic) scalar. This scalar aij k k x i y j means [see (ii)] that the various
k

aij are the components of a tensor.

So, with respect to a basis (e i e j e k ) , the n 3 expressions aij k are the components of a
tensor iff, for every arbitrary covector = k e k the n 3 expressions aij k k are the
components of a tensor.
The general tensor criterion is expressed as follows:
If the contracted multiplication of a mathematical entity and an arbitrary tensor is a tensor,
then the mathematical entity is a tensor.
Criterion. If the contracted multiplication (k times) of a mathematical entity and a tensor of
type qp leads to a tensor of type rs , then the mathematical entity is a tensor of type rs ++ kk qp .

()

()

In the previous example, the tensor (of components) aij k is really of type

0 +1 0
2 +11

(
) = ( ).

1
2

Example. The work done by a force f applied to a particle whose infinitesimal displacement
is represented by the vector dx is expressed as
dW = f i dx i .

The contracted multiplication of the entity of components f i and the vector of components
dx i leads to a scalar. So, the entity of components f i is a tensor of type

0 +11
0 +1 0

) = ( ).
0
1

So, the force is a covector.


Remark. The previous criteria are valid whenever the tensors are relating to the same basis
of a vector space. So, by considering the expression

dx p =

x p i
dx ,
x i

x p
are the components of a tensor of type
x i
since the various components dx p and dx i are not relating to the same basis.

it would be stupid to conclude that the various

1+11
0+10

)= ( )
1
1

164

Chapter 2

3. EUCLIDEAN VECTOR SPACE


Let E be an n-dimensional vector space that we are going to provide with a metric
from the definition of a new law: the scalar multiplication.
3.1

PRE-EUCLIDEAN VECTOR SPACE

3.1.1

Scalar Multiplication and Pre-Euclidean Vector Space

The scalar multiplication on E is a mapping


E E R : ( x , y )  x. y

such that, x , y, z E , k R :
P1.
P2.
P3.
P4.

x. y = y. x ,
( x + y ).z = x. y + y.z ,
x.( y + z ) = x. y + x.z ,
(kx ). y = k ( x. y ) = kx. y ,
[ y E : x. y = 0 ]

(commutativity)
(distributivity)
(mixed associativity)
x =0.

The real x. y is called the scalar product of x and y.


D

A vector space E provided with the scalar multiplication is called a pre-Euclidean


vector space.

We note that the first three properties show a symmetric bilinear form since it is linear with
respect to every element of any pair of vectors and is commutative. The fourth property shows
this form is nondegenerate.

3.1.2

Fundamental Tensor

From the preceding, we say:


PR8

The scalar multiplication on E defines the nondegenerate symmetric bilinear form


g : E E R : ( x , y )  g ( x , y ) = x. y ,

that is a tensor of type


D

( ).
0
2

The tensor g is called the fundamental tensor on E.


It is such that x , y E :

g ( x, y ) = x g y = x. y = x, y .

(2-29)

165

Tensors

Now, we express the tensor g by considering a basis (e i ) of E.


Given vectors x = x i e i and y = y j e j , the properties P2 and P3 imply:

x. y = x i y j e i .e j .

(2-30)

In a natural manner, we denote the components of g by




g (e i , e j ) = g ij .

(2-31a)

g ij = e i .e j = e i , e j

(2-31b)

So, we have:


and thus (2-30) is written:




g ( x, y ) = x. y = g ij x i y j .

(2-32)

The expression of g is thus




g = g ij e i e j .

(2-33)

Remark. The property P1 means that the scalar product is symmetric; that is, the tensor g is
symmetric:
g ij = g ji .

The property P4 is interpreted as follows:


[ y j : g ij x i y j = 0 ]

[ i : x i = 0 ]

or
[ g ij x i = 0 ]

[ i : x i = 0 ].

So, the system of n equations g ij x i = 0 only admits the trivial solution, and thus
g = det( g ij ) 0 .

Conversely, the reader will immediately verify that if n 2 reals g ij are such that g ij = g ji and
det( g ij ) 0 , then the scalar product fulfills the four properties of the scalar multiplication.

To conclude this section, we say:


D

Any two vectors x and y are orthogonal if their scalar product is zero:
x y

if

x, y = 0 .

166

Chapter 2

3.2

CANONICAL ISOMORPHISM AND CONJUGATE TENSOR

3.2.1

Canonical Isomorphism

Let us establish the canonical isomorphism existing between E and E . How?


Given a bilinear form
g : E E R : ( x, y)  g ( x, y ) ,

we express:
PR9 The bilinear form g being nondegenerate, there is a (canonical) isomorphism between
E and E defined by the flat mapping1:
b

: E E : x  xb = g ( x, )

such that, with every vector x is associated a covector defined by




g ( x , ) : E R : y  g ( x , y ) = x. y .

We observe that the real x. y is




g ( x, y) = xb ( y) = x , y

(2-34)

We note that the covector xb , that is g ( x , ) , is expressed with respect to a basis (e i ) of E


as
g ( x , ) = g ij x j e i

(2-35)

= g x
i

This last equality shows the contraction between g = g ij e e j and x = x k e k .


We really have v E :
g ( x, v ) = g ij x j e i . v k e k = g ij x j v k ki
= g ij x i v j .
In particular, with any basis vector e k E is associated the covector of E :


g (e k , ) = g ki e i .

Now we prove the proposition.


The flat mapping is linear since immediately x , y E :
( x + y ) b = g ( x + y, ) = g ( x, ) + g ( y, )
= xb + yb .

Also called lowering mapping (or application bmol in French).

(2-36)

167

Tensors

It is injective (one-to-one) since

xb = yb
( x y ) b = g ( x y, ) = 0
x=y

(since the bilinear form is nondegenerate).

It is surjective (onto) because we are going to prove that with every covector E is
associated one vector x such that = xb .
Indeed, we have
= j e j
where
j = (e j ) .
In addition, given the bilinear form of components g ij , the vector which fits the question has
its components xi such that
g ij x i = j

( xb = ).

( )

Therefore, there is a (unique) solution since the matrix g ij is nonsingular.


The proposition is so proved.
The isomorphism provides E with a pre-Euclidean structure in a natural manner.

3.2.2 Conjugate Tensor and Reciprocal Basis


D

The inverse mapping of a flat mapping is called a sharp or raising mapping1:


# : E E :  # .

Symbolically, we denote
(#)-1 = b .
Make explicit the components of any vector # dual of x by introducing the conjugate
b

tensor.

Under the previously defined isomorphism, to the components x i of x correspond the


following components of the covector xb :
x i = x b (e i ) = g ( x , e i ) = x j e j , e i
= g ij x .
j

( )

(2-37)

Denote by g ij the inverse matrix of (g ij ) ; it exists because det (g ij ) 0 . It is written without


rigor as
1

Called dise in French.

168

Chapter 2

g ij =
From equalities


cofactor of g ij
det( g ij )

g ij g jk = ki

(2-38)

g ij xi = g ij g ik x k = kj x k = x j .

(2-39)

we deduce

D  The g ij are the n 2 components of a tensor of type

( ) called the conjugate tensor of g.


2
0

It is denoted
g 1

and is expressed as


g 1 = g ij e i e j .

(2-40)

The bilinearity of g 1 is sure because this last is a function of covectors g ( x , ) and g ( y, )


such that:
g 1 ( g ( x , ), g ( y, )) = g ( x , y ) .
In an explicit manner, given
g 1 : E E R : ( g ( x , ), g ( y, ))  g 1 ( g ( x , ), g ( y, )) ,

we have
g 1 ( g ( x, ), g ( y, )) = g 1 ( xi e i , y j e j ) = g 1 (e i , e j ) xi y j
= g ij xi y j
and
g ( x, y ) = g rs x r y s = g rs g ir xi g js y j = g ir rj xi y j
= g ij xi y j .
Remark. The tensor character of g 1 is also revealed from a criterion of tensor calculus.

Indeed, the formula (2-39) expresses a contraction between the tensor of components g ij and

( ) . Since the contraction leads to a ( ) -tensor of


).
are the components of a tensor of type (

the tensor of components xi of type


components x j , we know that the g ij

0
1

1
0
1+1 0
0+11

The components of the conjugate tensor g 1 are




g ij = g 1 (e i , e j )

since
g rs e r e s (e i , e j ) = g rs ri sj = g ij .

(2-41)

Tensors

169

Given any covector E , we notice that the vector g 1 (, ) is expressed with respect to a
basis (e i ) of E as
g 1 (, ) = g ij j e i
= g 1
and in particular:

g 1 (e k , ) = g ki e i .

(2-42a)

We say:
D

The image of every vector of the dual basis (e i ) under g 1 ; that is, the following
vector of E
e k = g 1 ( e k , )
(2-42b)
is called the reciprocal vector of e k .

Thus, we have


e k = g ki e i

(2-42c)

e j = g kj e k .

(2-43)

and consequently


In addition, we have:
e i .e j = g ij ,

e i .e j = i j .

(2-44)

Indeed,
e i .e j = g iq e q .g jr e r = g iq g jr g qr = g iq qj = g ij ,
ei .e j = e i .g jk e k = g jk g ik = i j .

The reciprocal basis is the basis of E composed of vectors e i .

Of course, the vectors of a reciprocal basis are linearly independent; that is, the linear
combination
ck e k = 0
implies
k : ck = 0 .
Indeed, from
x E : 0 = (c k e k ) . x = c k x i ik = c k x k
we deduce
k : ck = 0 .

170

Chapter 2

We note that every vector of the reciprocal basis, for instance e k , is orthogonal to each of
n 1 vectors of the original basis (e i ) , except for i = k where the scalar product equals 1.
Classic example. In the usual 3-dimensional space, the reciprocal basis of a basis made up of
normed vectors (e i ) is composed of three vectors

ei =

e j ek
ei .(e j e k )

(i j k )

where ei .(e j e k ) is the volume of the parallelepiped constructed on the basis vectors.

3.2.3

Covariant and Contravariant Representations of Vectors

Fundamental remark

Given a basis (e i ) of E and the dual basis (e j ) , the formulae (2-37) and (2-38):

xi = g ij x j ,

x i = g ij x j

show the correspondence between any vector x E and its dual: the covector xb E .


The scalar multiplication allows identifying the isomorphic spaces E and E : the
reciprocal basis (e i ) and the dual basis (e i ) are identified.

In an explicit manner, we recall x E :


e k ( x ) = x k = x i ik = x i e i . e k = e k . x ,
which is also denoted:
x , e k = x , e k .
The fundamental remark leads to the following representation of every vector x of E.
D

The covariant representation of some vector x is the expression


x = xi e i
where the components

xi = x . e i
are called the covariant components of x.
This terminology follows from the fact that the components xi are the components of the
covector xb in the dual basis (e j ) of E since, given the covariant representation x = x j e j ,
we have:

Tensors

171

g ( x, ) = x j g (e j , ) = x j g ji g (e i , ) = x j g ji g ki e k
= x j e j .
In a general manner, the (usual) contravariant components x i are different from the covariant
ones as illustrated with the following:
Example. Let us consider a basis composed of normed vectors e1 and e 2 .

Fig. 63
The determination of the vector e1 of the reciprocal basis follows from the following
conditions:
e1 e 2
and
e 1 . e1 = 1 .
[likewise for e 2 ].
The covariant components of x are
x1 = x.e1 = x cos ,
x 2 = x.e 2 = x cos ,

while the contravariant components are the well-known x1 and x 2 .


The reader will immediately verify, on the figure, the necessary equality:
x1e1 + x 2 e 2 = x1e 1 + x 2 e 2 .

We conclude:


With the scalar multiplication, we can consider x and xb = g ( x, ) as a unique tensor


of order 1 for which exist two representations x i e i and xi e i .

172

Chapter 2

Contracted product and scalar product

In this new vision, we consider again the contracted product of any two tensors of
order 1; namely x , y E :
x yb = x i yi = x i g ij y j = x j y j = xb y .

(2-45)

It is nothing else but the scalar product


x. y = g ij x i y j
since
x. y = ( x i e i ).( y j e j ) = g ij x i y j = x j y j = x i yi .

3.2.4

Representations of Tensors of Order 2 and Contracted Products

Given the canonical isomorphism between E and E , we know that E E , E E ,


E E and E E are isomorphic and so are E E , E E , E E and E E .

(i) First, in particular, given vectors u and w, we consider the possible tensor products of
corresponding vectors and covectors:
t1 = ub w b ,

t 2 = ub w ,

t 3 = u wb ,

t4 = u w

where
ub = g (u, )

and

w b = g (w , ) .

We are going to make explicit these particular tensors of order 2; that is, x , y E :
x t1 y = t1 ( x, y ) = ub w b ( x, y ) = u i w j e i e j ( x, y )

= u i w j x i y j = ub , x w b , y = (ub x )( w b y ) ,
x t 2 yb = t 2 ( x, yb ) = ub w ( x, yb ) = ui w j e i e j ( x, yb )
= u i w j x i y j = ub , x w , yb = (ub x )( w yb ) ,
xb t 3 y = t 3 ( xb , y) = u w b ( xb , y) = u i w j ei e j ( xb , y )
= u i w j xi y j = u, xb w b , y = (u xb )( w b y ) ,
x b t 4 yb = t 4 ( x b , y b ) = u w ( x b , yb ) = u i w j e i e j ( x b , y b )
= u i w j xi y j = u, xb w , yb = (u xb )( w yb ).
Since we have ub x = u xb and w b y = w yb , the four previous reals are equal and we
denote the common value by
t ( x, y) = x t y

that is

173

Tensors
x (ub w b ) y = (ub x )( w b y )

(2-46)

where u and x are tensors of order 1 of opposite variances obviously and so are w and y.
Now, if we consider a tensor of type

( ), for example u w
1
1

E E , we have y E :

(u w b ) y = w b , y u

(2-47)

since
(u i w j e i e j ) y k e k = u i e i wk y k = w b , y u ,
[likewise for ub w E E ].
From the definition of the contracted multiplication and if we consider vectors and covectors
a, b, c, d such that the following operations are possible, then the reader will immediately
prove:
(a b) (c d ) = (b.c ) a d

(2-48)

and the double contraction:


(a b) : (c d ) = (b.c )(a.d ) .

(2-49)

(ii) Second, we view the general tensors of order 2 by considering the canonical isomorphism
between E and E . The generalized expressions of the previous tensors t1, t2, t3 and t4 are
obviously:

t1 = t ij e i e j E E ,
t 2 = t i e i e j E E ,
j

t 3 = t i j ei e j E E ,
t 4 = t ij e i e j E E .
By referring to the contravariant-contravariant representation of a general tensor of order 2,
namely:
t = t ij e i e j ,
we obtain the other representations as follows.
From
t = t ij e i e j = t ij e i g jk e k = t i k e i e k ,
we deduce the contravariant-covariant representation:
t = t i k ei e k

[with t i k = g jk t ij ],

the covariant-contravariant representation:


j

t = ti e i e j

[with t i j = g ik t kj ],

174

Chapter 2

and the covariant-covariant representation:


t = t ij e i e j

[with t ij = g ik g jl t kl ].1

The four previous representations are the ones of the second-order tensor t.
We conclude this section with the respective representations of the transposed tensor t t :
t

t = ( t t ) ij e i e j = t ji e i e j ,

t = ( t t ) i j ei e j = t j ei e j ,

t = (t t )i e i e j = t j i e i e j ,

t = ( t t ) ij e i e j = t ji e i e j .

So, the various representations of a symmetric tensor t = t t are such that t ij = t ji , t ij = t ji


and t i j = t j i .
We note that the symmetry of any tensor t of order 2 is expressed as follows:
x , y E :

xt y = yt x ,

since
x t y = ( t t x ). y = y.(t x ) ;

that is, for example:

t ij x i y j = t ji x i y j = t ij y i x j .

3.3

EUCLIDEAN VECTOR SPACES

Let E be a pre-Euclidean vector space.


D

The vector space E is said to be Euclidean if the symmetric bilinear form g defining
the scalar multiplication is positive-definite;
that is, for every nonzero x E :
g ( x, x ) > o .

(2-50a)

()
q

We specify that repeated contractions with g allow making canonically correspond tensors of types 0 and

( ), so we have :
0
q

t i1...iq = g i1 j1 ...g iq jq t
i ...iq

t1

= g i1 j1 ...g

iq jq

j1... jq

t j1... jq .

175

Tensors

In other words, by Euclidean vector space we mean pre-Euclidean vector space such that,
given any basis (e i ) :

x, x = g ij x i x j > 0 .

x 0 :

(2-50b)

The previous definition can be considered from the following signature notion.
Given a basis change e j = ij e i , we know that the formula of component change of a ( 02 )
tensor is
g rs = ri sj g ij .
If (g) designates the matrix ( g ij ) , we denote these equalities in the matrix context as follows:
( g ) = t ( g )

where t = ( ri ) is the transpose of = ( sj ) .


In order to present ( g ) as a diagonal matrix, we view as = OD where O is some
orthogonal matrix and D some diagonal matrix:

.
d nn

d11

So,
t

= tD tO = D O 1

( g ) = D O 1 ( g ) O D .

We can transform the matrix (g) into diagonal form by a good choice of orthogonal matrix O;
that is:

g11

1
.
= O ( g) O .

g nn

Thus we have:
d11

( g ) =

g11

d nn

d11

g nn

2
g11d11

.
=

d nn

.
2

g nn d nn

If we choose d ii = g ii 2 for every i, then ( g ) is a diagonal matrix with elements + 1 or 1,


the nonsingularity preventing some zero.
Note that the choice of d i i cannot alter the signs of diagonal elements.
For instance, the matrix O can be selected such that the +1 are in front.
Thus, the fundamental tensor can be written:
g = e 1 e 1 + ... + e p e p e p+1 e p+ 1 ... e p +q e p +q

with p + q = n .

176

Chapter 2

Several definitions of the signature of g have been given. Let us make the following choice.
D

The signature of g is the pair ( p, q ) of natural numbers such that p + q = n .

This is independent of the basis choice.


If q = 0 then the fundamental tensor is positive-definite, the signature shows positive signs
only. It characterizes every Euclidean vector space.
D

If the signature shows positive and negative signs, then it is said to be indefinite and
the corresponding pre-Euclidean vector space is called a pseudo-Euclidean vector
space.

If the fundamental tensor g is positive-definite, it is called the metric tensor of the


Euclidean vector space.

The norm of any vector x of a Euclidean vector space is the real

x =

x, x .

We note that if the space is pseudo-Euclidean, then

(2-51)

x, x is not a norm in the usual sense

because it can be a positive real, an imaginary number or zero, and

x, x

is said to be a

pseudo-norm.
D

A basis of a Euclidean vector space is said to be orthonormal if the basis vectors are
normed and orthogonal; that is:
ei .e j = ij ,

(2-52)

or in an equivalent manner:
g ij = ij

or
( g ij ) = I .

PR10 The covariant and contravariant components of every vector x with respect to any
orthonormal basis of a Euclidean vector space are such that
xi = x i ,
the variance being without significance.
Proof. We have:
x j = x . e j = ( x i e i ) . e j = g ij x i = ij x i = x j .

177

Tensors

Notation. The metric tensor appearing not any more, the Einstein summation convention is
not applicable any longer and
- either it is necessary to introduce the summation sign:
x

= (xi ) 2 ,
i

- or we adapt the Einstein convention, namely: If an index is repeated once, then it is a


dummy index indicating a summation with the index running through the appropriate integers
( i = 1,..., n ).
For example:
t ij xi x j .
Remark 1. In every orthonormal basis, the various representations of any tensor coincide.
Indeed, it is obvious that
e i = ei
( g ij = ij = g ij = g ij )
and for instance:
k
t ij l = t ijkl = t ijkl =
Remark 2. In the same manner, the formulae of transformation of components are for
example:
t i1...iq = i1 j1 ... iq jq t j1... jq .
Example. The metric tensor g and its conjugate g 1 are immediately expressed in relation to
the cylindrical coordinate basis.
Indeed, since any position vector is written
= r cos 1x + r sin 1 y + z 1z

with respect to the Cartesian orthonormal basis (1x ,1 y ,1z ) of the usual 3-dimensional space,
we deduce the vectors of the cylindrical basis, namely:

= (cos , sin , 0) ,
r

e =
= ( r sin , r cos , 0) ,

ez =
= (0 , 0 ,1) .
z
er =

We note this cylindrical basis is not normed (but orthogonal!) and the only nonzero g ij are

g11 = e r .e r = 1 ,

g 22 = e .e = r 2 ,

g 33 = e z .e z = 1 .

The matrices of the metric tensor and its conjugate are respectively:
1 0 0

(g ) = 0 r 2 0 ,
0 0 1

(g )
1

1 0 0

= 0 r 2 0 .
0 0 1

178

Chapter 2

4. EXTERIOR ALGEBRA
The works of H. Poincar and E. Cartan were at the root of exterior calculus. This
theory has been essential to develop the 20th century physics. We are going to consider
completely antisymmetric tensors.

4.1

p-FORMS
Let E be an n-dimensional vector space,
t be a p-linear form defined on p E .

4.1.1

Definition of a p-Form

A p-linear form is called skew-symmetric or completely antisymmetric if, for any


permutation making (1,..., p ) ( (1),..., ( p )) , we have x1 ,..., x p E :
t ( x (1) ,..., x ( p ) ) = t ( x1 ,..., x p )

(2-53)

where (or sign ) is +1 or 1 according to being an even or an odd permutation.


D

The alternation mapping or antisymmetrization is a (linear) mapping


A p : T p0 T p0 : t a A p t
such that x1 ,..., x p E :
A p t ( x1 ,..., x p ) =

1
t ( x (1) ,..., x ( p ) )
p!

(2-54)

where
is the permutation (1,..., p) ( (1),..., ( p)) ,
is the sum over all the permutations of the sequence (1,..., p) .

Remark 1. This sum is also denoted

S p

where S p is the symmetric group (order p!).

We note the presence of the conventional factor

1
and the sum over all p! elements of S p .
p!

Remark 2. The linearity of the antisymmetrization is obvious.


D

A p-form (or exterior form of degree p) is the image of a p-linear form by


antisymmetrization.
In other words:
A p-form is a completely antisymmetric tensor of type ( 0p ).

179

Tensors
Notation. The p-form A p t being well-defined, we denote it by a Greek letter:
= Ap t .

(2-55)

The set of p-forms is a vector subspace of T p0 denoted by p ( E ) or simply by p . Thus we


write:
p .

Example 1.

t T20 , x , y E :

( x, y ) = A2 t ( x, y ) = 12 (t ( x, y ) t ( y, x )) .
Example 2.

t T20 , e i , e j E (basis vectors):


(e i , e j ) = A2 t (e i , e j ) = 12 (t ij t ji ).

We evidently have the skew-symmetry property, namely:


(e i , e j ) = (e j , e i )

also denoted

ij = ji
Example 3.

t T30 , x1 , x 2 , x 3 E :
( x1 , x 2 , x 3 ) = A3 t ( x1 , x 2 , x 3 ) = 31! [t ( x1 , x 2 , x 3 ) + t ( x 2 , x 3 , x1 ) + t ( x 3 , x1 , x 2 )
t ( x1 , x 3 , x 2 ) t ( x 3 , x 2, x1 ) t ( x 2 , x1 , x 3 )] .

4.1.2

Exterior Product of 1-Forms

Given the 1-forms e i making up a basis of T10 , we have the successive and parallel
equalities:
A2 (e i e j )( x , y )

A p (e i1 ... e

= 12 (e i e j ( x , y ) e i e j ( y, x ))

1
p!

e i

i p

)( x1 ,..., x p )

... e

i p

( x (1) ,..., x ( p ) )

= 12 (e i e j e j e i )( x , y )
1 i1 ...i p I1

e
p! I1 ... I p

I p

= 12 IJij e I e J ( x , y )

where the symbols IJij are:

where the symbols I11 ...Ipp are:

0 if (IJ) is not a permutation of (ij)

0 if ( I 1 ...I p ) is not a permutation of (i1 ...i p ) ,

1 if (IJ) is an even permutation of (ij)

1 if ( I 1 ...I p ) is an even permutation of (i1 ...i p ),

-1 if (IJ) is an odd permutation of (ij).

-1 if ( I 1 ...I p ) is an odd permutation of (i1 ...i p ).

... e

( x1,..., x p )

i ...i

180

Tensors

In conclusion, we have:

A p (e i1 e

i p

i ...i
1
1 p e I1
p! I1...I p

)=

I p

(2-56a)

Notation. From now on, according to convention, we denote the 1-forms e i making up a
basis of T10 by i .

So the previous expression is written:


i

A p ( i1 ... p ) =

i ...i
1
1 p
p! I1... I p

Ip

I1 ...

(2-56b)

Remark.

IJij I J ( x, y ) = ( i j j i )( x, y )
= i ( x) j ( y) j ( x) i ( y)

D

xi

xj

= 2 A2 ( i j )( x , y ) .

The exterior product of p linear forms ik is the p-form 1


i1 ...

We note that the factor

1
p!

ip

i ...i

= I11...Ipp I1 ...

Ip

(2-57)

is not present in this definition. So we have conventionally chosen:

i1 ...

ip

= p! A p ( i1 ... p ) .

Several other conventions exist, but as the reader will see in the next, we have adopted the one
that eliminates the most constants.

Examples.

3 1 = 3 1 1 3 = 1 3 ,
ijk
i j k = IJK
I J K

= i j k + j k i + k i j
j i k i k j k j i .
These examples show the importance of the order of exterior product terms.
Exercise 21 proves that the exterior product of two 1-forms = i i and = j
= ( i j j i ) i j =
i< j

= i j i j .
1

The symbol can be called wedge or hat .

1
( i j
2!

j i ) i

is

181

Tensors
4.1.3

Expression of a p-Form

(i)

First, we consider the case of any 2-form 2 .

PR11 The C 2n products i j ( i < j ) form a basis of the vector space 2 .


Proof. On the one hand, every 2-form is a linear combination of C 2n =

n!
2! (n 2)!

products of

1-forms i j ( i < j ).
Indeed, let us express the value of on vectors x and y of E, more precisely the image of
(x,y) by ; namely:
( x , y ) = ( x i e i , y j e j ) = x i y j (e i , e j )
= ij x i y j

[letting ij = (e i , e j ) = A2 t (e i , e j ) ]
= 12 x1 y 2 + 21 x 2 y 1 + 13 x1 y 3 + 31 x 3 y 1 + ...
[since ii = 0 ]
= ( x1 y 2 x 2 y 1 ) + ( x1 y 3 x 3 y 1 ) + ...
[since ij = ji ]
12

13

= ij ( x y x y )
i

i< j

= ij ( i j j i )( x , y )
i< j

= ij i j ( x , y ) .
i< j

So, the products i j ( i < j ) generate every 2-form.


On the other hand, these products are linearly independent, namely:

ij i j = 0

i, j { 1,..., n }: ij = 0

i< j

because r, s { 1,..., n }:
0 = ij i j (e r , e s ) = ij ( i (e r ) j (e s ) j (e r ) i (e s ))
i< j

= 2 rs

i< j

(r < s) .

Therefore, the expression of a 2-form relative to the basis ( i j ) i< j is


= ij i

( ij = ji ) .

i< j

This sum contains C 2n =

n!
2! (n 2)!

terms.

The components of a 2-form relative to the basis ( i j ) i< j are called the strict
components and are denoted (ij ) .
So the expression of a 2-form relative to ( i j ) i< j is

= (ij ) i

by knowing that the sum is only over i < j .

182

Tensors

Of course, we have

= 21! ij i j .
Example. Let E be a 3-dimensional vector space.
The exterior product of two 1-forms , 1 is a 2-form with C 32 strict components,
namely:
(12 ) = 1 2 2 1 ,
(13) = 1 3 3 1 ,
( 23) = 2 3 3 2 ,
these being the components relative to the basis ( i j ) i< j of 2 .
(ii)

Second, let us view the following.

Example. We consider any 3-form 3 .


We recall that a tensor of type

( ) =
0
3

ijk

j k is completely antisymmetric if

ijk = jik = kji = jki = ikj = kij .


The value of for x , y, z E is:

( x, y, z ) = x i y j z k(e i , e j , e k ) = ijk x i y j z k
= 123 x1 y 2 z 3 + 231 x 2 y 3 z 1 + 312 x 3 y 1 z 2
+ 132 x1 y 3 z 2 + 321 x 3 y 2 z 1 + 213 x 2 y 1 z 3
+
and the antisymmetry of various ijk implies:

( x, y, z ) = 123 ( x1 y 2 z 3 + x 2 y 3 z 1 + x 3 y 1 z 2 x1 y 3 z 2 x 3 y 2 z 1 x 2 y 1 z 3 )
+
=

xi

xj

xk

ijk y i

yj

yk

zi

zj

zk

i < j <k

where
x i y j z k = i ( x ) j ( y ) k ( z ) = ( i j k )( x , y, z )

and so on.
Therefore, the expression of is immediately:

ijk i j k

i< j <k

by letting
i j k = i j k + j k i + k i

i k j k j i j i k .

(i < j < k )

183

Tensors
The strict components lead to
= (ijk ) i j k .

(iii)

Finally, we consider the case of any p-form.

A p-form is decomposable if there exists p 1-forms i such that


= 1 ... p .

PR12 The C np decomposable forms i1 ...

ip

(i1 < ... < i p ) make up a basis of the vector

space p .
Proof. This proposition can be proved as before or in a rather different manner as shown in
Exercise 22.
We specify that:
i ...i

( i1 ... p )(e j1 ,..., e j p ) = I11...Ipp ( I1 ...

Ip

i ...i

)(e j1 ,..., e j p ) = I11...Ipp Ij11 ... j pp

i ...i

= j11... jpp .

As previously, we can obtain the expression of a p-form, that is




...

i1 < <i p

i1...i p i1 ...

ip

(2-58a)

where i1 ,..., i p { 1,..., n }.


We also have:


p! i1...i p

i1 ...

ip

(2-58b)

It is essential to distinguish the n p components of a tensor of type ( 0p ) from the C np


components of the p-form obtained from antisymmetrization. Consequently, we define:
D

The components of a p-form relative to a basis ( i1 ... p ) i1< <i p of p are


called the strict components and denoted by
(i ...i )
(i1 < ... < i p ) .
1

So the expression of a p-form is




= (i1...i p ) i1 ...

ip

by knowing that the sum is only over i1 < ... < i p .

(2-59)

184

Tensors
i

Remark 1. Expression of a p-form relative to the basis ( i1 ... p ) of T p0 .


Let
i
t = t i ...i i1 ... p
1

be a tensor of

T p0 .

We have:
= A p t = A p t i1...i p i1 ...

ip

= t i1...i p A p ( i1 ... p ) = t i1...i p


=

1
t
p! i1...i p

i1 ...

ip

i ...i
1
1 p I1
p! I1...I p

...

Ip

The next to last equality allows to write:

j ... j = p1! j ... j t i ...i .


i1 ...i p

(2-60)

In particular, we find again

ij = ( A2 t ) ij = 12 (t ij t ji ) .
Remark 2. According to usage we call 0-form any scalar; that is an element of 0 (of

dimension C 0n = 1 ).
PR13 If the degree of a p-form is higher than the dimension n of the vector space E, then the
form is necessarily zero, that is p > n : p = {0}.
Proof. If p > n then two indices are necessarily equal and the p-form is thus zero.

4.1.4 Exterior Product of p-Forms


D


The exterior product of a p-form and a q-form is a ( p + q ) form defined by


=

( p + q )!
p! q!

A p + q ( ) ,

(2-61)

that is an element of the vector space p + q having C np + q dimensions.


Problem. What is the image of p + q vectors x1 ,..., x p+ q by the exterior product ?

From the definition of antisymmetrization, we have x1 ,..., x p + q E :


( )( x1 ,..., x p + q ) =

1
p!q!

( )( x (1) ,..., x ( p+q ) )

S p + q

185

Tensors
=

1
p!q!

1
p!q!

( x (1) ,..., x ( p ) ) ( x ( p +1) ,..., x ( p+q ) )

S p + q

1... p +q

i1...i p + q

i1...i p + q

( x i1 ,..., x i p ) ( x i p +1 ,..., x i p + q ) .

Remark. In particular, we find again the definition of exterior product of two 1-forms:

i j = 12!1!! A2 ( i j ) =

2!
( i
2!

j j i ) .

4.1.5 Exterior Algebra

Let us show that there are vector spaces whose direct sum is an algebra of finite
dimension.
Let us recall that the following (internal) law can be defined in the set of exterior forms of
same type.
D

The addition of two p-forms is the law


p p p : (, )  +

where + is the p-form sum,


the sum of two p-forms with respective (strict) components (i1 ...i p ) and (i1 ...i p ) being
the p-form with the C np following components:
(i1 ...i p ) + (i1 ...i p ) .
We can also define a second law:
D

The multiplication of a p-form by a scalar k is an (external) law:


R p p : (k ,)  k

where k is the product of by k,


the product of the p-form with (strict) components (i1 ...i p ) by the scalar k being the
p-form with the C np components:
k ( i1 ...i p ) .

The exterior product space of vector spaces p and q is the vector space of
(p+q)-forms; that is,
p q = p +q .

186

Tensors

Now we can define a third law that is bilinear.


D

The exterior multiplication of a p-form and a q-form is the mapping


: p q p + q : (, )  ,

the exterior product being such that x1 ,..., x p + q E :


( x1 ,..., x p + q ) =

1... p+q

i1...i p + q

1
p!q!

i1...i p + q

( x i1 ,..., x i p ) ( x i p +1 ,..., x i p + q ) .

The following properties are verified:

(i)

The exterior multiplication is bilinear:


p , , q : ( + ) = +
, p , q : ( + ) = +
k R , p , q : k ( ) = k = k .

Proof. The first equality (the others being demonstrated in the same way) is immediate:
( + ) = 2! A2 ( ( + )) = 2!( + )
= 2! A2 ( ) + 2! A2 ( ) = + .
(ii)

The exterior multiplication is associative:


p , q , r :
( ) = ( ) =

(2-62)

Proof.
[( ) ]( x1 ,..., x p + q , x p + q +1 ,..., x p + q + r )
=

1
( p + q )!r !

1... p +q +r ( x i ,..., xi
i1...i p + q + r

i1...i p + q + r

p+q

) ( x i p + q +1 ,..., x i p + q + r )

p + q+ r n

but
( x i1 ,..., x i p + q ) =

1
p!q!

j1... j p + q

j ... j

i11...i p +pq+ q ( x j1 ,..., x j p ) ( x j p +1 ,..., x j p + q )

thus
[( ) ]( x1 ,..., x p + q + r )
=

1
( p + q )! r ! p! q!

i1...i p + q i p + q +1...i p + q + r j1... j p + q

j ... j

i ...i

...i

i11...i p +pq+ q 11... p +p +qq... p +pq++q +r r

. ( x j1 ,..., x j p ) ( x j p +1 ,..., x j p + q ) ( x i p + q +1 ,..., x i p + q + r ) .


At this level we recall the notion of signature of a permutation.
We can view the permutation (i1 .....i m ) ( j1 ..... j m ) as the composite of two other
permutations:

187

Tensors

i1 ....im k1 ....k m i1 ....im

=

,
j1 ... j m j1 ... j m k1 ...k m
but the signature of the composite of two permutations being the product of their signatures,
then we have:
ij11......ijmm = ki11......ikmm kj11......jkmm .
So the permutations imply
j1... j p + q i p + q +1 ...... i p + q + r i1....i p + q i p + q +1 ....... i p + q + r j1.. j p + q i p + q +1...i p + q + r

1... p + q p + q + 1.. p + q + r = 1... p + q p + q + 1... p + q + r  i ...i i

...
i
1 p + q p + q +1 p + q + r

and
j .. j

...i

j ... j

..i

i ..i

...i

p + q p + q +1 p + q + r
1...1 p +pq+ qp +pq++q1+...1 p +pq++q +r r = i11....i pp++qqi pp++qq++11...i pp++qq++rr 11..........
............. p + q + r

j .. j

i ...i

...i

p + q p + q +1 p + q + r
.
= i11...i pp++qq 11..........
............ p + q + r

Therefore, the sum relative to the indices i1 ...i p + q , in [( ) ]( x1 ,..., x p + q + r ) is such that:

j1 .. j p + q
i1 ...i p + q

1...................... p + q + r =
i1 ..i p + q i p + q +1 ..i p + q + r

i1 ..i p + q

j1 .. j p + q i p + q +1 ....i p + q + r
1... p + q p + q +1... p + q + r

1
p + q p + q +1
p+q+r
= ( p + q )! 1..........
.................. p + q + r .

j ... j

...i

i1 ..i p + q

Finally,
[( ) ]( x1 ,..., x p + q + r )

1
r ! p!q!

j1.. j p + qi p + q +1... p + q + r

1......................... p +q +r

j1.. j p + q i p + q +1..i p + q + r

( x j1 ,..., x j p ) ( x j p +1 ,.., x j p + q ) ( x i p + q +1 ,.., x i p + q + r ).

In the same way, we can show this expression is


[ ( )]( x1 ,..., x p + q + r ) ,

which proves the associative property.


(iii)

The exterior multiplication is not commutative and p , q :


= (1) pq .

(2-63)

Proof. On the one hand, we have:


( )( x1 ,..., x p , x p +1 ,..., x p + q )
=

1
p!q!

1................... p+q ( xi ,..., xi


i1...i p i p +1...i p + q

i1...i p + q

) ( x i p +1 ,..., x i p + q )

and on the other hand:


( )( x1 ,..., x p , x p +1 ,..., x p + q )
=

1
p!q!

1.................. p+q ( xi
i p +1...i p + q i1...i p

i1...i p + q

p +1

,..., x i p + q ) ( x i1 ,..., x i p ).

188

Tensors

The permutation (i p +1 ...i p + q i1 ...i p ) (i1 ...i p i p +1 ...i p + q ) corresponds to pq transpositions; thus
we have:
= (1) pq .
Special cases
If p and if p is odd, then = 0 .

If , 1 , then = and = 0 .
D

The exterior algebra is the vector space


n

( E ) = 0 1 ... n = i ,
i =0

direct sum of vector spaces which is necessarily finite.


This last precision follows from the fact there is no exterior form of degree p higher than n.
The previous vector space is provided with a bilinear law: the exterior multiplication. We can
express:
PR14 The exterior algebra is associative, noncommutative and of finite dimension.
The exterior algebra is a vector subspace of the tensor algebra T whose dimension is
dim 0 + ... + dim n = C 0n + C1n + ... + C nn = 2 n .
Let us specify that the product of an element of degree p with an element of degree q is of
degree p + q (graduation!).

4.2 q-VECTORS
The previous developments about p-forms are applicable to q-vectors by replacing E
by E and thus the reader will easily transpose the main notions about p-forms into the
context of q-vectors.
D

A q-vector is the image of a q-linear form (defined on q E ) by antisymmetrization.


In other words:
A q-vector is a completely antisymmetric tensor of type

( ).
q
0

The exterior product of q vectors eik is the q-vector


I ... I

ei1 ... e iq = i11...iq q e I1 e I q .

(2-64)

The C qn decomposable q-vectors ei1 ... e iq ( i1 < < iq ) make up a basis of the vector
space of q-vectors.

189

Tensors

The expression of any q-vector is a q-linear form:

w=

i ...iq

w1

i1 <<iq

e i1 ... e iq

where i1 < < i q {1,..., n} and w 1

= w (e i1 ,..., e

i ...iq

iq

).

The components of a q-vector relative to the previous basis are called the strict
components w

( i1...iq )

We denote
w=w

(i1...iq )

e i1 ... e iq .

(2-65)

Example 1. Let us consider any 2-vector w. We have successively:

w = w ij ei e j = w e i e j + w ij e i e j
ij

i< j

i j

= w ij e i e j + w ji e j e i
i< j

i j

= w ij e i e j w ij e j e i
i< j

(since w ij = w ji , w ii = 0)

i< j

= w ij (e i e j e j e i )
i< j

= w (ij ) e i e j .
In particular, the exterior product of two vectors x, y E :
x y = x y y x = ( x i y j x j y i )e i e j
is such that
w ij = x i y j x j y i

and thus
x y=
i< j

xi

xj

yi

yj

ei e j .

This completely antisymmetric tensor has C 2n strict components


( i < j ).

xi y j x j y i

More particularly, concerning the usual 3-dimensional space, the C 32 = 3 strict components
are
x1 y 2 x 2 y 1 ,
x1 y 3 x 3 y 1 ,
x 2 y 3 x3 y 2 .
Example 2. Let us consider the exterior product of three vectors x , y, z E . We have:
xi
x yz=

i< j <k

xj

xk

yi y j

y k ei e j e k .

zi

zk

zj

190

Chapter 2

This particular 3-vector has C 3n strict components represented by the determinants.


More particularly, concerning the usual 3-dimensional space E, the C 33 = 1 strict component is
x1

x2

x3

y1 y 2

y3 .

z1

z3

z2

It is the well-known mixed product:


x. y z = x y.z
defining the volume (with sign) of parallelepiped having x , y, z as sides.
Example 3. Let us consider the exterior product of q vectors:
x i1
t = x y ... z =

y i1

i1 <<iq

z i1

...

xq
i

... y q
e i1 ... e iq ,
....
i

... z q

where x , y,..., z are q vectors of an n-dimensional vector space E.


It is a completely antisymmetric tensor of type

( ) having C
q
0

q
n

strict components.

In particular, if the number of vectors is equal to the dimension n of the space E, then there is
only one strict component and we have clearly:

t = t (12...n ) e1 e 2 ... e n .
So, the components of t relative to a basis of the space of tensors of type

( ) are such that


n
0

t i1...in = i1...in t (1...n )


where

=0

i1...in = 1

if two indices are equal,


if the number of transpositions with respect to (i1 ...in ) is even,

= 1 if the number of transpositions with resp e ct to (i1 ...in ) is odd .

More particularly, concerning a 3-dimensional space, the components of a completely


antisymmetric tensor of type 30 are such that

()

t ijk = ijk t
where t is the strict component.

191

Tensors
Basis change

The reader having carried on with the comparison between p-forms and q-vectors can
now conclude this section with the notion of basis change.
For example, given a 2-dimensional space E, we know that a tensor of type

( ) is such that
2
0

t = t (12) e1 e 2 = t (12) ( 11e1 + 12 e 2 ) ( 21 e1 + 22 e 2 )


= t (12) ( 11 21 12 21 ) e1 e 2

but it is also
t = t (12) e1 e 2

and thus

(12 )

=t

(12)

11 12
21

22

If E is an n-dimensional space, we have:


t (ij ) e i e j = t (ij ) ( ip e p ) ( qj e q )
= t (ij ) ( i1e1 + + in e n ) ( 1j e1 + + nj e n )
= t (ij ) [( i1 2j i2 1j ) e1 e 2 + ( i1 3j i3 1j ) e1 e3 + ]
= t (ij ) [ ir sj is rj ] e r e s

but it is also

t ( rs ) e r e s
and thus
t

( rs )

=t

(ij )

ir

is

rj

sj

In general, given a p-vector t, a change of basis of an n-dimensional space E, leads to

ir is ... in
t ( rs...n ) = t (ij...k )

rj sj ... nj

(2-66)

. . . . .

kr ks ... kn
In the following paragraph, we are going to consider affine spaces also called point spaces
that were recalled in chapter 0.

192

Chapter 2

5. POINT SPACES
In mechanics, the consideration of the point space associated with a vector space is
essential, it allows to introduce the notion of tensor fields for instance.

5.1

POINT SPACE AND NATURAL FRAME

Let E be a (real) vector space of dimension n, that we will suppose to be preEuclidean.


5.1.1 Point Space

We know that there exists an n-dimensional point space E associated with E and a
mapping
E E E : (a, b)  ab
such that:
(i)
(ii)
(iii)

a, b E :
a, b, c E :
o E , v E , ! x E :

ab = ba ,
ab + bc = ac ,
ox = v .

Remark 1. It follows from (i) and (ii) that


ab = ao + ob ,

that is
ab = ob oa .

Remark 2. Given an arbitrary point o E , condition (iii) establishes a bijection between E


and E; that is, given o the knowledge of point x implies the knowledge of v and conversely.

5.1.2 Coordinate System and Frame of Reference

Let us recall the following notions.


D

A frame of E is the set { o; e i } composed of a point o E and vectors of a basis (e i ).


The point o is the origin of the frame.

The coordinates of any point x E relative to a frame { o; e i } are the components of


ox E with respect to the basis ( ei ).

Therefore, the coordinates x i of x E are such that

ox = x i e i .

193

Tensors

We say that the coordinates x i ( i = 1,..., n) form a system of coordinates ( x i ).


Given points a and b of respective coordinates a i and b i , from

ab = ob oa = (b i a i ) e i
we deduce that the components of ab E relative to the basis ( ei ) are the differences
between coordinates of respective points b and a.
Change of coordinate system

Let { o; e i } and { o ; e j } be frames of reference.


Let us establish the relations between the coordinates of a given point x E with respect to
both frames.
Given

oo = Ai e i ,

oo = B i e i ,

ox = x i e i ,

ox = x i e i ,

and since the relations between basis vectors are

ei = i j e j ,

e j = ij ei ,

then we have:

ox = oo + ox = Ai e i + x i i j e j = ( Ai + x' j ij ) e i .
Therefore, we deduce by identification that
x i = Ai + ij x j .

In the same manner, by expressing o x with respect to the basis (e i ) , we immediately obtain
x i = B i + ij x j .
Coordinate systems
D

The correspondence between the points of E and n-tuples of reals u i (called the
coordinates) defines a coordinate system on E.
Every coordinate line is the locus of points of which only one coordinate varies.

For instance, u 1 is varying while the other coordinates are invariable ( u 2 , u 3 ,... ).
Another coordinate curve can be defined by u 2 = k (k constant) and so on.
So, we have a mesh of n coordinate curves defining a coordinate system.
We know that any straight line of E is a one-dimensional subspace of E determined for
instance by a fixed point a E and a fixed vector w E ; it is defined by
ax = k w

where k is a real parameter.

194

Chapter 2

Given a fixed frame of reference { o; ei0 } of E , we know that the straight line equation is

ox = oa + ax .
From oa = a i e i0 and w = w i e i0 , we deduce that the straight line equation is

ox = (a i + k w i ) e i0 .
The position vector of x being ox = x i ei0 , the straight line is defined by the n following
equations:
x i = a i + k wi .
The coordinates x i of the running point x being first degree functions of the parameter k, we
say:
D

The coordinate lines being straight, the coordinates are called rectilinear coordinates
with respect to a given fixed frame.

Now, given a fixed frame { o; ei0 } of E , we consider the rectilinear coordinates x i of x E as


arbitrary functions of n arbitrary real coordinates u j , namely:
x i = i (u j )

i, j = 1,..., n .

We assume the following:


The functions i are (at least) of class C 2 , the Jacobian such that
J=

( x1 ,..., x n )
(u 1 ,..., u n )

on a given domain of (u j ) , and we can express:


u j = j (xi )

i, j = 1,..., n.

In general, with these new coordinates u j the coordinate lines are not straight lines
obviously. So, we say:
D

The coordinates corresponding to (general case) coordinate curves are called


curvilinear coordinates.

Remark. The curvilinear coordinates u j are rectilinear iff the rectilinear coordinates x i are
x i
j
are constant.
first degree functions of u ; that is iff the various
u j

5.1.3 Natural Frame

We consider the rectilinear coordinates x i of a point x E relative to a fixed frame


{ o; ei0 }; in other words, we define:

ox = x j e 0j .

195

Tensors

We know that there are n coordinate lines trough the point x and let u i be the curvilinear
coordinates.
The Jacobian being different from zero, we denote the tangent vectors to different coordinate
lines by
ox
.
(2-67)
ei =
u i
Their expressions with respect to the basis (e i0 ) are
ei =

x k 0
ek .
u i

We note that the derivatives of ox do not depend on point o since, for every point o E
different from o, we have:
o x = ox oo = ox + c

and thus

o x
u

ox
u i

(constant vector c)

So, we write naturally:


ei = i x .

(2-68)

These n vectors which are tangent to n coordinate lines through x are linearly independent
since the Jacobian J is nonzero and we say:
D

The n vectors e i = i x constitute a basis called a natural basis or a (natural)


coordinate basis.

A frame composed of any point x E and n basis vectors through x is called a natural
frame.

It is denoted
{ x, i x }.
D

The differential of ox is the vector


dox =

x
du i ,
i
u

(2-69a)

also denoted

dx = e i du i

(2-69b)

where the (contravariant) components of the differential vector are du i .


Example 1. Let us show the local frame at point x of the classical Euclidean plane in the case
of polar coordinates (differential presentation).

Let x1 , x 2 be the (Cartesian) rectilinear coordinates,


r , be the (curvilinear) polar coordinates.

196

Chapter 2

A natural frame is shown in the following figure.

Fig. 64
The Cartesian frame being { o; e10 , e 20 ], we write:
dx = dx1 e10 + dx 2 e 20
= (cos dr r sin ) e10 + (sin dr + r cos d ) e 20

but
dx = r x dr + x d
and thus
e1 = cos e10 + sin e 20 ,
e 2 = r sin e10 + r cos e 20 .
So, the natural frame is made up of the point x and two orthogonal vectors e1 and e 2 tangent
to respective coordinate lines corresponding to constant values of on the one hand and to
constant values of r on the other hand. Their respective norms are 1 and r.
Example 2. Let us show the natural frame at point x of the classical 3-dimensional Euclidean
space in the case of spherical coordinates (derivative presentation).

We know that the Cartesian coordinates and the (curvilinear) spherical coordinates are related
by
x1 = r sin cos ,
x 2 = r sin sin ,
x 3 = r cos
where r, and are respectively the radial distance, the colatitude and the longitude.
Given the Cartesian frame { o; e10 , e 20 , e 30 }, the various vectors of a natural basis are:

197

Tensors
e1 = 1 x = sin cos e10 + sin sin e 20 + cos e 30 ,
e 2 = 2 x = r cos cos e10 + r cos sin e 20 r sin e 30 ,
e3 = 3 x = r sin sin e10 + r sin cos e 20 .

Fig. 65
The natural basis vector e1 has the radial direction of ox ( and being constant), e 2 is
tangent to the local meridian ( and r being constant) and e3 is tangent to the local parallel (r
and being constant).
The orientation senses are induced by the ones of r, and respectively.
So, the natural frame is made up of the point x and three orthogonal vectors of respective
norms 1, r and r sin .

5.2

TENSOR FIELDS AND METRIC ELEMENT

We are going to define one tensor at each point of the point space E , but first we
consider curvilinear coordinate transformations.
5.2.1

Transformations of Curvilinear Coordinates

We consider a transformation of curvilinear coordinates


R n R n : (u i )  (u j )

such that the functions f j , defined by


u 1 = f 1 (u 1 ,...u n )

. . .

u n = f n (u 1 ,..., u n ) ,

198

Chapter 2

are of class C q ( q 2 ).
The Jacobian

(u 1 ,..., u n )
is assumed non-vanishing.
(u 1 ,..., u n )

Conversely, we have:
u j = g j (u k )

j , k = 1,..., n .

Whenever there is such a coordinate transformation, there is a corresponding change of frame


at a given point.
More explicitly, we consider a point x and a correspondence between two natural bases of
respective vectors
ei =

x
,
u i

e j =

u j

(2-70)

Let us seek the formulae of transformation of tensor components given the previous
coordinate transformation simply denoted
u j = u j (u i )

i, j = 1,..., n .

First, we say:
PR15 Natural frames { x; e i } and { x; e j } are respectively associated with any systems of
curvilinear coordinates (u i ) and (u j ) such that the formulae of change of natural
bases are
u i
(2-71)
e j =
ei ,
u j
u j
(2-72)
ei =
e j .
u i
By remembering previous notations of basis changes, we note that in this case we have:

ij =

u i
,
u j

ij =

u j
.
u i

Proof. From the differentiable composite mapping theorem, we have obviously:


x
u i
x
ei =
u j
e j =

which proves the proposition.

u i
,
u j
u j
,
u i

199

Tensors

Remark. We specify that the fixed Cartesian frame { o; ei0 }, from which the various ei have
been introduced, does not appear in this context of curvilinear coordinate transformations. So,
the correspondence between natural bases associated with different coordinate systems is only
a correspondence between curvilinear coordinate systems at a given point.

PR16 The set of natural bases at x E is identical to the set of bases of the vector space E
associated with the point space E.

Proof. At a given point x E , by considering the set of frames { x; e i } which correspond to


various curvilinear coordinate systems, we know that each natural basis (e i ) is a basis of the
vector space E associated with E.
Conversely, any basis (e j ) of E is a natural basis at x E by considering the formulae of
coordinate transformations as

e j =

u i
ei .
u j

Any tensor of the vector space E is said to be a tensor on the point space E (associated
with E).

This is well-defined from the previous proposition. Any tensor t on E can be expressed with
respect to a local basis (e i ) associated with the curvilinear coordinate system (u i ) .
Transformation of tensor components

We are going to find the formulae of transformation of tensor components under a


curvilinear coordinate transformation at x E .
For a tensor t of type

( ) , we have:
1
0

t = t j e j = t i e i = t i

u j
e j
u i

and thus

t j =
For a tensor t of type

u j i
t ,
u i

ti =

u i j
t .
u j

(2-73)

ti =

u j
t j .
u i

(2-74)

t ij =

u i u j pq
t
u p u q

( ) , we obtain in the same manner:


0
1

t j =

u i
ti ,
u j

Likewise, we have:

t pq =

u p u q ij
t ,
u i u j

200

Chapter 2

u p u j i
t j,
=
u i u q

u i u q p
t q
=
u p u j

and so on.
For example:

t pqr s t =

u p u q u r u m u t ijk n
t m .
u i u j u k u s u n

So, we say:
Rule. Each position of index of primed components of any tensor at x E is clearly imposed:
u
u
To any upper index corresponds one
and to any lower index corresponds one
.
u
u

(Inversely for unprimed components).


Example. Gradient of a real-valued function.

We consider the function


f : R n R : (u 1 ,..., u n )  f (u 1 ,..., u n )

and a coordinate transformation defined in a simplified writing as


u j = u j (u i )

i, j = 1,..., n .

The differential
f
f
du j
du i =
i
j

u
u
j
f u
du i
=
j
i
u u

df =

implies

f
u j f
=
.
u i
u i u j
So, the various

f
are the components of a tensor of type
u i

This tensor is called the gradient of f.


Its components with respect to any frame { x, e i } are
Its Cartesian coordinates are the successive

f
.
x i

f
.
u i

( ).
0
1

201

Tensors
5.2.2

Tensor Fields

From previous developments, we are going to assign only one tensor at each point
xE .

If we denote the set of tensors of type


D  A tensor field of type

( ) at x E by T
q
p

q
xp,

we say:

( ) on E is a mapping
q
p

t : E ( {x} Tx p ) : x  t ( x ) = ( x, t x )
q

xE

where t x is a

( )-tensor assigned to x.
q
p

We know that, at each point x E , a tensor can be expressed with respect to a natural basis
(e i ) of E associated with a coordinate system (u i ) .
The operations on tensors assigned to a same point pose no problem since the tensors are
relative to natural frames at this point (for which changes of natural bases have been defined).
For example, the sum of two tensors is clearly defined from their components with respect to
different natural bases at the same point, because the tensors can be expressed with respect to
a same natural basis.
However, difficulties arise for tensors assigned to different points! How to compare tensors
related to natural frames at different origins? It is necessary to study the change of natural
frame between two infinitely neighboring points by referring to the natural frame at one of the
points.
We are going to consider pre-Euclidean spaces; more precisely, we recall:
D

A point space E is pre-Euclidean (resp. Euclidean) if the associated vector space E is


pre-Euclidean (resp. Euclidean).

5.2.3 Metric Element


D

The metric element of the point space E (also called line element) is the square of the
distance between two infinitely neighboring points of E.

Given any two infinitely neighboring points x and x + dx, we know that their distance is the
norm of vector defined by this pair of points, i.e. dx .
We denote the metric element by
2

ds 2 = dx .

202

Chapter 2

(i) First, we consider a rectilinear coordinate system ( x i ) on E.


The coordinates of points x and x + dx are respectively x i and x i + dx i with respect to the
rectilinear frame { o; eio }.
From

dx = dx i e io
we express the metric element as

ds 2 = dx

= dx i e io . dx j e oj

or
ds 2 = g ijo dx i dx j ,

(2-75)

where we have obviously let g ijo = e io .e oj .


Remark 1. The metric element of E and thus the components g ijo of the fundamental tensor g

with respect to the rectilinear basis (e io ) completely define the metric of E.


We also note that the various g ijo are not dependent on x, they are constant.
Remark 2. In relation to an orthonormal basis of a Euclidean space frame, the metric
element of the Euclidean space is immediately:

ds 2 = ij dx i dx j = dx i dx i = dxi dxi .
(ii) Second, we consider a system of curvilinear coordinates (u i ) such that:
x j = x j (u i )

i, j = 1,..., n .

Given a natural frame { x; e i } as before defined, the expression of the metric element of the
point space is
ds 2 = g ij du i du j
(2-76)
where the various g ij = e i .e j are the components of the fundamental tensor g.
Remark. The metric element of E and thus the components g ij of g with respect to the

natural basis (e i ) completely define the metric of E. But here, we note that the various g ij are
dependent on x and are obviously not constant.
A metric tensor is assigned at each point of E and is an example of tensor field.
Example. The metric element relative to spherical coordinates r , , such that
x = r sin cos ,

y = r sin sin ,

z = r cos

is the sum
ds 2 = dr 2 + r 2 d 2 + r 2 sin 2 d 2

of the following radial dr 2 , meridian (r d ) 2 and longitudinal (r sin d ) 2 squared


displacements.

Tensors
5.3

CHRISTOFFEL SYMBOLS

5.3.1

Definition of Christoffel Symbols

203

Given a coordinate system (u i ) on E, the essential problem is to determine, at any


point x + dx , a natural frame in comparison with the natural frame at an infinitely
neighboring point x.
Relating to a natural frame { x; i x }, we have:
dx = i x du i = e i du i .
The natural frame { x; e i } is deformed from x to x + dx and becomes { x + dx; e i + de i }. The
problem is to determine the increments de i of natural basis vectors.
The differentials of vectors of the natural frame { x; e i } are the following vectors expressed
with respect to the natural basis (e i ) :


de j = ij e i

(2-77)

where the n 2 matrix elements ij are such that


de j . e k = ij g ik = jk .
We will later give a physical interpretation of these elements.
Remark. From the relation between reciprocal basis vectors ei .e j = i j , we deduce:

d (e i . e j ) = de i . e j + de j . e i = 0

and thus
ei . de j = ik e k . e j = ij .

So, the (covariant) components of de j are ij and the covariant representation of vector
differentials is
de j = ij e i .
(2-78)
Now, we can introduce the Christoffel symbols.
The vectors of the natural basis (e i ) depending on coordinates u k , we have obviously:
de j = k e j du k .

(2-79)

From (2-77) and (2-79), we deduce that the vectors k e j are expressed with respect to the
natural basis (e i ) so much so that we say:

204

Chapter 2

The Christoffel symbols ijk are n 3 functions1 of curvilinear coordinates u k that are
the components of vectors k e j :

k e j = ijk e i .

(2-80)

i
ijk
= k e j , e i

(2-81)

de j = ijk du k e i

(2-82)

We also note that

We have immediately

and we point out the differential forms

ij = ijk du k .

(2-83)

PR17 The Christoffel symbols are not the components of a tensor (in general).
Proof. Given a curvilinear coordinate transformation u p = u p (u i ) , we have:
de p = d (

u i
u i
u i
de i
e
)
=
d
(
)
e
+
i
i
u p
u p
u p

but
de i = ikj du k e j

de p = prq du r e q

thus
u i
u i j
du k e j
+
e
)
i
p
p ik
u
u
i
q
i
q
u u
j u
k u

= d(
du
eq
) i e q + ik
p
p
j
u u
u
u
q
i
q
k
2 i
u
u
j u u u
r

=
du e q + ik
du r e q
p
r
i
p
j
r
u u u
u u u

prq du r e q = d (

and so
prq = ikj

u i u q u k
2 u i u q
+

u p u j u r u p u r u i

(2-84)

The second term of the right-hand member means that the Christoffel symbols are not the
components of a 12 -tensor.

()

Remark 1. The lateral position of the upper index with respect to the lower indices does not
matter. In addition, we note that the symbols ij are not tensors too.

Sometimes called Christoffel symbols of second kind and denoted

{ }.
i
jk

205

Tensors

Remark 2. We note that the Christoffel symbols define tensors if the affine (or linear)
changes of curvilinear coordinates are such that

2u i
=0
u p u r
for every i, p and r.
D

The Christoffel symbols of first kind are n 3 functions of curvilinear coordinates u k


such that
ijk = g ih hjk
(2-85)
where the first index is the descending one.1

We have obviously
ijk du k = g ih hjk du k = g ih hj = ij
that is

ijk du k = de i . e j .

(2-86)

ikj = g jh hik .

(2-87)

We can also write :

Property of symmetry. The lower indices are interchangeable.

Indeed, from the (necessarily supposed) equality of second derivative vectors


j ( i x ) = j e i = ijk e k
and
i ( j x ) = i e j = kji e k
we deduce
ijk = kji .
In the same manner, we have
kij = kji

and the last two indices are interchangeable.


Given this property of symmetry, these relations are n n

The Christoffel symbol of first kind is also denoted

There are other definitions such that ijk

[i, jk ] .

= g jh ikh = [ik , j ].

n +1
2

in number.

206

Chapter 2

5.3.2 Ricci Identities and Christoffel Formulae


(i)

From the n(n + 1) 2 functions g ij = e i . e j of curvilinear coordinates u k , we deduce:


dg ij = e i . de j + de i . e j = e i . hj e h + ih e h . e j

that is
dg ij = g ih hj + g jh ih .

So, by considering the coordinates u k , we obtain the Ricci identities


k g ij = g ih hjk + g jh ikh .

(2-88)

Since for any value k there are n(n + 1) 2 terms k g ij , then there are n 2 (n + 1) 2 Ricci
identities.
(ii)

The Christoffel formulae are obtained from the Ricci identities written as follows:
ijk + jik = k g ij

(1)

and, after cyclic permutations of indices and by using the symmetries, they are:
jik + kij = i g jk

(2)

kij + ijk = j g ik .

(3)

and

From (1) + (3) - (2), we obtain the first Christoffel formula:




ijk = 12 ( j g ik + k g ij i g jk )

(2-89)

The second Christoffel formula is


ijk = g ih hjk = 12 g ih ( j g hk + k g hj h g jk ) .

(2-90)

Examples. The Christoffel formulas let calculate the Christoffel symbols given a metric (and
thus given the various g ij ):

In polar coordinates:

212 = 221 = r ,
2
2
21
= 12
=

122 = r ,
1
22
= r ,

the other symbols being zero.


In spherical coordinates:
1
22
= r ,
1
33
= r sin 2 ,

the other symbols being zero.

2
2
12
= 21
=1 r ,
2
33
= sin cos ,

3
3
32
= 23
= cot ,
3
3
13
= 31
=1 r,

207

Tensors

In cylindrical coordinates:

221 = 212 = r ,

122 = r ,

2
2
12
= 21
=1 r ,

1
22
= r ,

the other symbols being zero.


Remark. As long as the tensor g does not exist (no pre-Euclidean structure), the only method
to calculate the Christoffel symbols consists in going back to rectilinear coordinates by
expressing
x i 0
ej =
ei
u j
and thus
2 xi
ke j =
e0 .
j
k i
u u

[See also Exercise 26].

5.4

ABSOLUTE DIFFERENTIAL, COVARIANT DERIVATIVES, GEODESIC

5.4.1 Absolute Differential of a Vector, Covariant Derivatives

We consider a vector field


v : E ( {x} T01 ) : x  v ( x) = ( x, v x ) .
xE

The field vector at point x is v i e i with respect to the basis of the natural frame { x; e i }.
From any x to a neighboring point x + dx , the components of any field vector (as well as the
basis) change.
The corresponding variation of v is:
dv = dv i e i + v i de i
= dv i e i + v i i j e j

and thus we say:


D  The absolute differential of a vector field is a vector field defined by

dv = (dv i + v k ki ) e i

(2-91)

208

Chapter 2

So, the corresponding components with respect to (e i ) are not dv i apart from when the
coordinates are rectilinear, since the various ijk and thus de i vanish.

The (contravariant) components (dv ) i of the absolute differential dv with respect to


(e i ) are denoted
v i = dv i + v k ki
= ( j v i + v k kji ) du j

and are also called the absolute differentials of contravariant components v i .


PR18 The various


j v i = j v i + kji v k

are the components of a tensor of type

(2-92a)

( ) called the ( )-covariant derivative of v.


1
1

1
1

Proof. Since the various du j are the components of an arbitrary vector with respect to (e i )
and since v i are the components of the vector dv with respect to this basis (e i ) , then a
tensor criteria relative to the contracted multiplication lets assert that the various components
j v i + ijk v k are the ones of a tensor of type 11 also denoted by

()

j v i = v i , j + kji v k .

(2-92b)

[See also Exercise 31].


Now, if we consider the reciprocal basis (e i ) and thus the covariant representation
v = vi e i ,

then we have the following absolute differential of v:

dv = dvi e i + vi de i
= (dv j vi ij ) e j
and we say:
D

The covariant components (dv ) j of the absolute differential dv with respect to (e i )


are denoted
v j = dv j vi ij
(2-93)
= ( k v j vi ijk ) du k
and are also called the absolute differentials of the covariant components v j .

209

Tensors

PR19 In a general manner, the various components


j vi ijk vk
are the ones of a tensor of type

( ) denoted by
0
2

j vi = j vi ijk v k

(2-94)

= vi , j ijk v k
and are called the covariant components of the

( ) -covariant derivative of v.
0
2

Proof. Given a transformation of coordinates


u j = u j (u i )

i, j = 1,..., n

we know that the primed components of v = vi e i are expressed in function of unprimed


components as follows:

v j =

u i
vi
u j

and thus

v j
u k

vi u i
2u i
vi
+
u k u j u k u j

but from

u p u r u q
2 u i u r
+
u j u i u k u j u k u i
i
p
2u i
u q
r u
i u
= jk
,
pq
u j u k
u r
u j u k

jkr = ipq

we deduce:

v j
u k

p
vi u m u i
u q
r
i u

vi
+

jk r
pq
u m u k u j
u j u k

that is

v j
u k

jkr v r = (

vi
u

s
mi
vs )
m

u m u i
.
u k u j

So, we have proved that j vi = j vi ijk v k is a tensor of type

( ).
0
2

5.4.2 Absolute Differential of a Tensor, Covariant Derivatives

We consider a tensor field


t :E

( {x} T pq ) : x  t ( x) = ( x, t x ) .

xE

210

Chapter 2

From any x to a neighboring point x + dx , the components of any field tensor (as well as the
basis) change.
We are going to determine the differential of any tensor of the field.
Without restricting the theory, we choose tensors of order 3, for instance in order to simplify
the notation:
t = t i j k ei e j e k .

Let us determine the components t i j k of the absolute differential tensor with respect to the
basis of the natural frame { x; e i }, this tensor being denoted

dt = t i j k e i e j e k .
Given a fixed basis (e i ) , for every vector v = v i e i and every one-form = i e i of class C 1
we have:

d ( (v )) = d ( i v i ) = d i v i + i dv i = d (v ) + (dv ) .
In particular, since e i (e j ) = ij , we have:

d (e i (e j )) = de i (e j ) + e i (de j ) = 0 .
From

de i (e j ) = e i (de j ) = e i ( kj e k ) = kj e i (e k ) = ij
we deduce

de i = ij e j .

(2-95)

Now, from (2-77) and (2-95), we can determine the components t i j k of dt . Indeed,
dt = dt i j e i e j e k + t i j d (e i e j e k )
k

= dt i j e i e j e k + t i j ( ih e h e j e k e i hj e h e k + e i e j kh e h )
k

= dt i j e i e j e k + t h j hi e i e j e k t i h e i hj e j e k + t i j e i e j hk e k
k

= (dt i j + hi t h j hj t i h + hk t i j ) e i e j e k
k

So, given a natural basis (e i ) , we say:


D

The absolute differential of a


components are

( ) -tensor field t is a ( ) -tensor field dt


2
1

2
1

t i j = dt i j + hi t h j hj t i h + hk t i j
k

h
h

i
k i
= ( p t i j + hp
t h j hjp t i h + hp
t j ) du p
k

= p t i j du p .

of which the

211

Tensors

PR20 The various


k

k i
i
q t i j = q t i j + hq
t h j hjq t i h + hq
t j

are the components of a tensor of type

()
2
2

(2-96)

called the covariant derivative of t.

Proof. Since the various du p are the components of an arbitrary vector with respect to (e i )
and since t i j k are the components of the tensor dt with respect to the same basis, then a
tensor criteria relative to the contracted multiplication lets us assert that the various
components q t i j k are the ones of a tensor of type 22 . The proposition is so proved.

()

Now, the reader can immediately write the covariant derivative of any tensor.
For example, the components of the covariant derivative of t = t ij e i e j are
i hj
q t ij = q t ij + hq
t + hqj t ih .

(2-97)

In the same manner, we have:


q t ij = q t ij iqh t hj qjh t ih .

(2-98)

More particularly, from the Ricci identities (2-88), we immediately deduce:


q g ij = 0 .

So, the following Ricci theorem is stated:


PR21 The absolute differential of the fundamental tensor g is zero:
dg = 0 .

5.4.3

(2-99)

Geodesic and Eulers Equations

Consider the following curve c on a Euclidean point space:


[a, b] E : t  c(t )

such that, given the subdivision a = t 0 < t1 < < t p = b of [a, b] , the continuous function
t  c(t ) is differentiable on each [t i , t i +1 ] for every i [0, p 1].

The curve being defined by the parametric equation u i = u i (t ), we say:


D

The length of arc c is the real


lc =

g ij

du i du j
dt

dt

dt .

(2-100a)

212

Chapter 2

The length of c is also denoted with the curvilinear parameter s as follows:


lc =

g ij du i du j =

sb

sa

ds .

(2-100b)

The geodesic equations can be found from the calculus of variations which consists in
obtaining extremal curves; in this context, the geodesic distance between two points is the
lower bound of lengths (if it exists) of curves joining the points, the geodesic arc is the arc
corresponding to the lower bound.
Let us note that a geodesic does not necessarily lead to a minimum of length. Furthermore, the
sphere example proves there are several geodesics joining two diametrically opposite points
(an infinity).
Now, let us recall a fundamental result of the variational calculus and let us consider (class
C 2 ) curves joining two points.
With each curve is associated a previously defined length:
l c = F (u i ,
c

du i
, t ) dt = ds
c
dt

where
F (u i ,

du i
, t ) = g ij u i u j
dt

i, j = 1,..., n .

A geodesic being a curve such that the variation

sb

ds = 0 ,

the variational calculus teaches us that a (necessary) condition of extremum of l c is obtained


if the system of Euler equations
d F
F
( i) i =0
dt u
u

is verified, that is
j
d g ij u
1
(
)
i ( g jk ) u j u k = 0 .
dt F
2F

Substituting the curvilinear parameter s for t, the geodesic equations are


1
du j 1
du j du k
d
( g ij
)
i ( g jk )
=0
dt ds 2 ds
dt dt
dt

g ij

du k du j 1
du j du k
d 2u j
+

g
=0
k ij
i jk
ds ds 2
ds ds
ds 2

(2-101)

213

Tensors

du k du j
du j du k
du j du k
d 2u j 1
1
+ ( k g ij
+ j g ik
) i g jk
= 0.
g ij
ds ds
ds ds
ds ds
2
2
ds 2

Multiplying by g ri , we obtain the following system of geodesic equations




j
k
d 2u r
r du du
+ jk
=0.
ds ds
ds 2

(2-102)

5.4.4 Parallel Transport


Given a vector field v, in order to compare two vectors v (x) and v ( x + dx) at
infinitesimal neighboring points x and x + dx , we introduce the notion of parallel transport.
D

A parallel translation or parallel transport of v is a translation of the vector


v ( x + dx) along a geodesic from x + dx to x.

Let v x ( x + dx) denote the corresponding parallel translated vector at x.


In rectilinear coordinates, from x + dx to x, the components of the parallel translated vector
do not vary since the local frames have the same basis.
On the other hand, in curvilinear coordinates, the parallel translation modifies the vector
components since the local frame basis changes.
Throughout this section, the reader will think about the well-known example of polar
coordinates. He will translate the vector of field v from x + dx to x. He will compare this
vector v x (x + dx) with the vector of field at x:
v x (x+ dx) v(x)

and he will express the components of this difference with respect to the local basis.
PR22 The absolute differential of v ( x) is
dv = v x ( x + dx) v ( x)
Proof. The components of the unit vector tangent to the geodesic being

(2-103)
du i
, we consider
ds

the following scalar product:


v . = vi

du i
.
ds

The increment of this scalar product between two infinitesimal neighboring points of the
geodesic is
du i
du i
d 2u i
) = dvi
+ vi
ds .
d (v . ) = d (vi
ds
ds
ds 2

214

Chapter 2

d 2u i
By expressing
from the geodesic equation, we have:
ds 2
d (v . ) =

vi

du j

du j du k
du i
vi ijk
ds
ds ds
ds

u j
v k
du j du k
i
= ( j vi jk )
ds
ds ds
u

that is:
d (v . ) = j v k du j

du k
ds

also denoted by
du k
ds
= dv . .

d (v . ) = (dv ) k

This last equality leads to a geometric interpretation of the absolute differential dv. Indeed,
the increment d (v. ) is the difference between the following scalar products

d (v. ) = v ( x + dx) . v ( x) .
= v x ( x + dx). v ( x).
= (v x ( x + dx) v ( x)).
and we deduce at x:
dv = v x ( x + dx) v ( x) .

5.4.5 Absolute Derivative of a Vector (Along a Curve)

Let (u i ) be a curvilinear coordinate system in a point space E.


Let x be a vector field such that to each value of any parameter t corresponds only one vector
of the field and we denote
x = x (t ) .

By considering
ei =

x
= i x ,
u i

we recall that the differential vector of x is

dx = i x du i = du i e i .
If the parameter t is the time in particular, then
by v.

dx
is called the velocity of x and is denoted
dt

215

Tensors

The components of the velocity v with respect to (e i ) are


du i
v =
.
dt
i

(2-104)

This generalizes, in curvilinear coordinates, the classical velocity relative to (Cartesian)


rectilinear coordinates.
Now, from the notion of absolute differential of a vector field v and by considering an
increment dt of parameter t making pass (along a curve) of point x to a neighboring point
x + dx, we define:
D

The absolute derivative of vector field v = v i e i along a curve is the vector field
dv
dv i v k i
=(
+ k ) ei
dt
dt
dt
i
dv
du j
=(
+ v k ijk
) ei .
dt
dt

If the parameter t is the time,

dv
is called the acceleration of x.
dt

The components of the acceleration with respect to (ei ) are


du j
dv i
+ ijk v k
dt
dt
j
k
2 i
d u
i du du
=
+ jk

dt dt
dt 2

ai =

(2-105)

Remark. Geodesics are curves followed by points of zero acceleration.

Indeed, if the acceleration vanishes then, with a system of rectilinear coordinates ( x i ) , we


have:
d 2 xi
i
a =0=
i = 1,..., n
dt 2

xi = k it + K i

( k i , K i : constants),

and for a system of curvilinear coordinates (u i ) we have:


j
k
d 2u i
i du du
+

=0.
jk
dt dt
dt 2

216

Chapter 2

5.5

VOLUME FORM AND ADJOINT REPLACE WITH NEW

Let us consider a coordinate transformation defined by


u j = u j (u i )

i, j = 1,..., n .

5.5.1 Volume Form

First we consider the example of a completely skew-symmetric tensor of type

( ).
0
3

In the case of a 3-dimensional space, we recall that such a tensor has one strict component; let
us denote this by D.
The components of the

( ) -tensor relative to the basis ( e


0
3

e 2 e 3 ) are

Dijk = ijk D

where the Levi-Civita symbol is

ijk

0 if at least two indices are equal


= 1 if (ijk ) is an even permutation of (123)

1 if (ijk ) is an odd permutation of (123).


Every change of component is clearly denoted by

D pqr =

u i u j u k
ijk D .
u p u q u r

Since

=(
D123

u 1 u 2 u 3
+ ) D ,
u 1 u 2 u 3

the change of strict component is thus


u i
D = det
j
u

D.

(2-106)

This formula can be generalized to completely antisymmetric tensors of type


dimensional spaces.

()
0
n

on n-

The physicists call D a tensor density.


The components of a completely antisymmetric tensor of type
are
t i1...in = i1...in D .

( ) relative to (e
0
n

e n )

Exercise. The reader will immediately verify that a completely antisymmetric tensor C of
type 0n on an n-dimensional space shows the following change of strict component

()

C = (det(

u i 1
)) C .
u j

217

Tensors

He will prove that the components of such a tensor relative to (e i1 e in ) are


t i1...in = i1...in C .

Now, let us define the volume form on a Euclidean space.


Under any basis change defined by
e j = ij e i

u i
ei
u j

or

e j =

or

=
g kr

we have:
= ki rj g ij
g kr

u i u j
g ij .
u k u r

By denoting
det g = det( g ij ) ,

the rule of product of determinants is applied, namely:


det g = (det ) 2 det g ,

det g = (det(

u i 2
)) det g .
u j

We assume det g and det g are positive; so, we consider direct basis changes (on the
oriented space). We have:
det g ' = det
So,

u i
det g ' = det
j
u

det g .

det g is a tensor density, it is the strict component of a completely antisymmetric tensor

of type
D

det g ,

( ). Hence, we express:
0
n

The volume form on the Euclidean space is the following completely antisymmetric
n-form
= det g e 1 ... e n .

(2-107)

The components of relative to the basis (e i1 e in ) are

i1...in = i1...in det g .


Remark 1. The volume spanned by vectors e1 ,, e n is

(e1 ,..., e n ) = det g .


In particular, in 3-dimensional Euclidean space, we have:

ijk = (e i , e j , e k ) = e i e j . e k

(2-108)

218

Chapter 2

and is the tensor associated with the trilinear mapping making a correspondence from any
( x , y, z ) to the mixed product ( x y. z ) .
In this example, we note that the various ijk are the components of a tensor since we
consider bases of same orientation only; but it would not be a tensor any more for bases of
different orientations.
Remark 2. The following change of strict component
1

det g

= (det(

u i 1
))
u j

lets the introduction of a completely antisymmetric tensor of type


and having the following components relative to (e i1 e in )
1

i1...in = i1...in

det g

(2-109)

det g

( ), dual of volume form


n
0

(2-110)

Note that

i1...in i1...in = n!
since the previous sum contains n! equal terms as

12...n 12...n = 213...n 213...n = = 1 .

5.5.2 Adjoint

From the p-form notion, the duality has allowed us introducing the q-vector notion.
The reader has transposed the developments about p-forms; for example he has shown that the
q-vectors are elements of a C qn -dimensional vector space.
Let t be a completely antisymmetric tensor of type
i ...iq

The corresponding q-vector of components t 1


i ...iq

t1

= g i1 j1 ...g

iq jq

i ...iq

The adjoint of the q-vector t 1


denoted t such that
(t ) iq +1...in =

q!

( ).
0
q

is such that

t j1... jq .

(2-111)

relative to the volume form is the (n q ) -form


i ...iq

i1...in t 1

So, the operator defines an (n q ) -form from a q-vector and conversely.

(2-112)

219

Tensors

We also denote the following contravariant components:


(t )

i p +1...in

p!

i1...in t i1...i p .

In the same manner, we say:


D

The adjoint of the p-form i1...i p relative to is the (n p ) -vector, denoted , such
that
()

i p +1...in

1 i1...in

i1...i p .
p!

(2-113)

Remark 1. The adjoint of a completely antisymmetric tensor of type


scalar.
Remark 2. The adjoint of a completely antisymmetric tensor of type
covector [resp. vector].

( ) or of type ( ) is a
0
n

( )
n 1
0

n
0

[ resp.

( ) ] is a
0
n 1

This notion lets find again mathematical notions of elementary geometry on R 3 as curl and
divergence. For instance, we consider the following
Example. Given an orthonormal basis of classical 3-dimensional Euclidean space, the vector
product of two vectors is the adjoint of the exterior product of two vectors.
Indeed, let us consider
t = x y = x i y j ei e j = ( x i y j x j y i ) ei e j
= ( x i y j x j y j ) ei e j .
i< j

The adjoint of this

( ) -tensor is the covector of components


2
0

(t ) k =

i j
1
2! ijk ( x y

x j yi )

= 12 det g ijk ( x i y j x j y i )
= det g (ij ) k ( x i y j x j y i ).

or
(t ) k = det g ijk x i y j

(2-114)

that is
(t )1 = det g ( x 2 y 3 x 3 y 2 ) ,
(t ) 2 = det g ( x 3 y 1 x1 y 3 ) ,

(2-115)

(t ) 3 = det g ( x y x y ) .
1

2 1

By considering an orthonormal basis of Euclidean space ( det g = 1 ) we know that the


variance is indifferent and we find again the components of the vector product x y :
x 2 y 3 x3 y 2 ,

x3 y1 x1 y3 ,

x1 y 2 x 2 y1 .

220

5.6

Chapter 2
DIFFERENTIAL OPERATORS

We are going to give various presentations of differential operators.


5.6.1

Gradient

5.6.1a Gradient of a Function

Let f : R n R be a differentiable real-valued function of curvilinear coordinates u i .


We are going to consider the differential of f denoted
df =

where the various

f
du i
i
u

f
behave like components of a covector field called the gradient of f.
u i

Let us recall the notions of differential and directional derivatives of f.


D

The differential of f at point x is the linear mapping

df x : R n R : h  df x (h)
such that

f ( x + h) f ( x ) = df x (h) + h (h)
lim (h) = 0 .

h0

Let us make explicit this mapping.


Let (e i ) be the natural basis.
If we consider the special vector h1 e1 of R n , we have:
f (u 1 + h1 , u 2 ,..., u n ) f (u 1 , u 2 ,..., u n ) = h1df x (e1 ) + h1 e1 (h1e1 ) .

But the following


lim (h1e1 ) = 0

h1 0

implies
lim
1

h 0

f (u 1 + h1 ,..., u n ) f (u 1 ,..., u n )
= df x (e1 ) .
h1

Consequently, at x, the function f possesses a partial derivative with respect to u 1 :


df x (e1 ) =

f
( x) .
u 1

221

Tensors

Generally, we have:
df x (e j ) =

f
( x) .
u j

(2-116)

Now, if we consider any vector h = h j e j of R n , the differential of f, at x, is the (linear)


mapping
df x : R n R : h  df x (h) = h1 df x (e1 ) + + h n df x (e n )
and we write:
df x (h) =

f
( x) h j
j
u

j = 1,..., n

(2-117)

We also recall the following:


D

The directional derivative of f, at x, in the direction of a (unit) vector v is the limit (if
existence):
Dv f ( x) = lim

k 0

f ( x + kv ) f ( x )
.
k

(2-118)

In particular, the partial derivative with respect to u j is the directional derivative in the
direction of the vector e j of the natural basis (e i ) :
De j f ( x) =

f
( x) .
u j

Now, we say:
D

The gradient of a function f : R n R , at x, is the covector


grad x f : R n R : h  grad x f (h)
such that
grad x f (h) = df x (h)
that is explicitly:
( grad x f ) i h i =

The gradient of f is also denoted


x f .

f
( x) h j .
j
u

(2-119)

222

Chapter 2

So, the components of grad x f with respect to the natural basis (e i ) are:
( grad x f ) i = i f ( x)

(2-120)

also denoted
i f (x) .
In conclusion, the covector


grad x f = i f ( x) e i

(2-121)

is such that
grad x f . h = i f ( x) e i . h j e j
= i f ( x) h i .
The corresponding gradient field is the covector field denoted by
grad f .

In differential writing, by letting


h = dx = du i e i ,
we have1:
grad x f . dx = ( grad x f ) i du i = i f ( x) du i
= df x .

(2-122)

Remark 1. The gradient of a function is obviously determined from the knowledge of the
tensor g.

We recall that the natural basis associated with curvilinear coordinates is made up of vectors
x
ei = i , where x is the position vector of a point x; each ei is tangent to the u i coordinate
u
curve at x.
By denoting ei the unit vector in each tangent direction, we write:
ei = k (i ) ei .

From the differential vector dx = e i du i also written:


dx = k (1) du 1 e1 + k ( 2) du 2 e 2 + k (3) du 3 e3 ,
we deduce for orthogonal curvilinear coordinates
ds 2 = dx . dx = k (21) (du 1 ) 2 + k (22) (du 2 ) 2 + k (23) (du 3 ) 2 .
1

For a nice and complete presentation see section 2.3 of lecture 0 of our book Differential Geometry with
Applications to Mechanics and Physics .

223

Tensors

Therefore, the gradient of f , which is


grad x f =

k (1)

f
1 f
1 f
e +
e +
e3 ,
1 1
2 2
k ( 2) u
k (3) u 3
u

(2-123)

is easily determined by knowing that the various k (i ) are such that:


k (2i ) = g ii .
For example, in spherical coordinates r , , , the gradient field is expressed in the
orthonormal basis (e r , e , e ) as
grad f =

f
f
1 f
1
e r +
e +
e .
r
r
r sin

Remark 2. By considering orthonormal bases with curvilinear coordinates, Eq. (2-122) is


simply denoted (for any point):

df = i f du i = i f du i
or
df = i f e i .

(2-124)

So, given spherical coordinates r , , , we have with respect to the orthonormal basis
(e r , e , e ) :
dr

df = ( r f , f ,
f ) r d = r f dr + f d + f d
r
r sin
r sin d

and in this case, the 1-forms e i are


(e ) r = dr ,

(e ) = r d ,

(e ) = r sin d .

In the same manner, given cylindrical coordinates r , , z , we have with respect to the
orthonormal basis (e r , e , e z ) :
dr

df = ( r f , f , z f ) r d = r f dr + f d + z f dz
r
dz

and in this case, the 1-forms e i are


(e ) r = dr ,

(e ) = r d ,

We are going to generalize the previous notions.

(e ) z = dz .

224

Chapter 2

5.6.1b Gradient of a Tensor

Let t (x) be a tensor field whose tensor components are differentiable with respect to
the curvilinear coordinates u i of the position vector x of x E .
D

The directional derivative of a tensor field t (x) in the direction of a (unit) vector v is
the tensor field such that x E :
t ( x + kv ) t (v )
.
k

Dv t ( x) = lim

k 0

(2-125)

In particular, the partial derivative with respect to u j is the directional derivative in the
direction of the vector e j of the natural basis (e i ) :
De j t ( x) =

t
( x) = j t ( x) .
u j

(2-126)

In conformity with the gradient of a function, we say:


D

The gradient of a tensor field t (x) of type


that:
grad t ( x) = j t ( x) e j

( ) is the tensor field of type ( ) such


q
p +1

q
p

(2-127)

= De j t ( x) e j

such that h = h i e i E :

grad t ( x) . h = j t ( x) e j . h i e i = j t ( x) h j .
Remark 1. The symbol does not appear in (2-121) where the tensor is of type

( ).
0
0

Remark 2. We note that the linear directional derivative operator in the direction of h,
namely:
Dhi e = h i Dei
i

lets us deduce the following:


Dh t ( x) = grad t ( x) . h

since
Dh t ( x) = Dhi e t ( x) = h i i t ( x) = grad t ( x) . h
i

In differential writing, we have:


grad x t . dx = d t x .

(2-128)

Tensors

237

Exercise 8.
Prove that the sum of
tensor of the same type

( ).

( ) -tensors
2
1

t = t ij k e i e j e k and u = u ij k e i e j e k is a

2
1

Answer. From
t ij k = ip qj kr t pq r ,
u ij k = ip qj kr u pq r
we deduce:
(t ij k + u ij k ) = ip qj kr (t pq r + u pq r )
and thus t + u is really a tensor of type

( ).
2
1

Exercise 9.
Given tensors of components i j and t pq r , express the contraction in indices i and p.

Answer. Since i j are components of the Kronecker tensor, the contraction leads to a tensor
of type

( ) of components
2
1

i j t iq r = t jq r .

Exercise 10.
Prove that the contraction of the tensor product of t = t p e p and u = u q e q is a tensor
of type

( ).
0
0

Answer. By considering
t u = t p u q e p e q ,
a change of basis leads to

t i u j = ip qj t p u q
and the contraction implies:

t i ui = ip iq t p u q = pq t p u q
= t pu p .

238

Chapter 2

Exercise 11.

By contracting a tensor of components t pq rst in p and t, and next in q and s, prove that
the result is a covector.
Answer. First, by summing on the repeated index p = t , we have:
t ij kli = mi nj ka lb ic t mn abc
= mc nj ka lb t mn abc
= nj ka lb t mn abm

and so we have obtained a tensor of type

( ) , denoted by
1
2

u n ab ( = t mn abm ).
Second, by summing on the repeated index n = b , we have:
u j kj = nj ka bj u n ab = nb ka u n ab

= ka u b ab .
We have really obtained a covector.
Exercise 12.

Given two tensors t = t ij e i e j and u = u kl e k e l , prove that


t : u = tr (t u) = u : t

and
t : I = tr (t ) .

Answer. The contracted product t u is the following second-order tensor

t u = t ij u jl e i e l .
The only supplementary contraction leads to the scalar

t : u = t ij u ji ;
but the trace of the tensor t u is obviously the scalar following from the contraction in the
indices i and l; that is:
tr (t u) = t ij u ji = t : u .

We have immediately:
u : t = u ij t ji = t ij u ji
= t :u

In addition, we have:
t : I = tr (t I ) = t ij ji = t ii .

239

Tensors
Exercise 13.

Given a change of basis of a Euclidean vector space

e1 = 2e1 + e 2 ,

e 2 = e1 + 2e 2 ,

express the components g ij in function of various g ij .


Answer. The change of basis e j = ij e i is such that

11 = 2 ,

12 = 1 ,

12 = 1 ,

22 = 2

and thus

= 1111 g11 + 11 12 g12 + 1211 g 21 + 1212 g 22


g11
= 4 g11 4 g12 + g 22 ,
= g 21
= 2 g11 3 g12 + 2 g 22 ,
g12
= g11 + 4 g12 + 4 g 22 .
g 22
These results can be verified from the definition g ij = e i .e j ; for instance:
= (2e1 + e 2 ).(e1 + 2e 2 )
g12
= 2 g11 + g 21 4 g12 + 2 g 22 .
Exercise 14.

Prove that , E :
g 1 = g 1 (, ) g g 1 ( , ) .

Answer. We have:
g 1 (, ) g g 1 ( , )
= ( g ij j e i ) ( g rs e r e s ) ( g pq q e p )
= g ij j g is g sq q = iq g ij j q = g qj j q
= j g js e s q e q
= ( j e j g rs e r e s ) q e q
= g 1 .
Exercise 15.

Given a vector a = a i e i and a basis (e k ) such that


e1 = e1 ,

e 2 = e1 + e 2 ,

e3 = e 2 + e 3

(i) Find the contravariant and covariant components of a with respect to (e k ) .

240

Chapter 2

(ii) Given a = e1 + e 2 + e 3 and b = 2e1 + e 2 e 3 , from the contravariant and covariant


components of a and b with respect to (e k ) , calculate the scalar product a . b .
Answer. (i) From
a = A i ei = A1e1 + A 2 (e1 + e 2 ) + A 3 (e 2 + e 3 )
= ( A1 + A 2 ) e1 + ( A 2 + A 3 ) e 2 + A 3 e 3

we deduce
a1 = A1 + A 2 ,

a 2 = A 2 + A3 ,

a 3 = A3

and thus the contravariant of a with respect to (e k ) are


A1 = a1 a 2 + a 3 ,

A2 = a 2 a 3 ,

A3 = a 3 .

The covariant components of a with respect to (e k ) are


A1 = a . e1 = a i e i . e1 = a1 ,
A2 = a . e 2 = a i e i .(e1 + e 2 ) = a1 + a 2 ,
A3 = a 2 + a 3 .

They are also obtained as follows:


A j = a . e j = Ai ei . e j = g ij Ai

with
g11 = 1 ,
g 22 = 2 ,

g12 = g 21 = 1 ,
g 23 = g 32 = 1 ,

g13 = g 31 = 0 ,
g 33 = 2 .

So, we find again


A1 = g1i Ai = A1 + A 2 = a1 ,
A2 = g 2i Ai = A1 + 2 A 2 + A3 = a1 + a 2 ,
A3 = g 3i Ai = A 2 + 2 A3 = a 2 + a 3 .
(ii) From
A1 = 1 ,
B1 = 0 ,

A2 = 0 ,
B2 = 2 ,

A3 = 1 ,
B 3 = 1

A1 = 1 ,
B1 = 2 ,

A2 = 2 ,
B2 = 3 ,

A3 = 2 ,
B3 = 0 ,

and

we deduce
a . b = A i Bi = Ai B i = 2 .

Of course this result can be found immediately.

241

Tensors
Exercise 16.

Given a two-dimensional Euclidean vector space, let t = t i jk e i e j e k be a tensor


of type

( ) such that
2
1

t111 = 1 ,

t112 = 1 ,

t121 = 1 ,

t122 = 2 ,

t 211 = 1 ,

t 212 = 2 ,

t 2 21 = 0 ,

t 2 22 = 1 .

(i) Calculate the contraction in first and second indices.


(ii) Given the fundamental tensor

(g ij )

3
=
2

determine the components t ijk .


(iii) Calculate the various components t ij k .
Answer. (i) The contraction leads to a tensor of type

( ) of components u
1
0

= t i ik which are:

u 1 = t111 + t 2 21 = 1 ,
u 2 = t112 + t 2 22 = 0 .

So, the tensor of type

( ) is e .
1
0

(ii) Since the conjugate tensor is

(g ) = 12
ij

2
,
1

we have:
t 111 = g 11 t111 + g 12 t 211 = 3 ,
t 112 = g 11 t112 + g 12 t 212 = 5

and so on:
t 121 = 1 ,

t 122 = 0 ,

t 211 = 3 ,

t 212 = 4 ,

t 221 = 2 ,

t 221 = 2 3 ,

t 22 2 = 5 3 .

(iii) From
k

t ij = g jp t i

pk

we deduce:
t111 = g11 t111 + g12 t1 21 = 1 3 ,
t112 = g11 t112 + g12 t122 = 1 ,
t121 = g 21 t111 + g 22 t121 = 1 3 ,
t12 2 = g 21 t112 + g 22 t122 = 0

and so on:
t 211 = 1 3 ,

t 212 = 4 3 ,

t 222 = 3 .

242

Chapter 2

Exercise 17.

Given two vectors x = x i e i and y = y j e j , find the expression of the double


contraction ( x y ) : (e k e l ) .
Answer. From
x y = x i y j ei e j ,

we deduce:
( x y ) : ( e k e l ) = x i y j (e i e j ) : (e k e l )
= x i y j (e j .e k )(e i .e l )
= g jk g il x i y j
= xl y k
= ( x.e l )( y.e k ).
In particular, if x = e i and y = e j , we have:
(e i e j ) : (e k e l ) = (e i .e l )(e j .e k ) = g il g jk .

Exercise 18.

Prove that

( )

det cij = ijk c1i c 2 j c3k .


Answer. We have:

ijk c1i c2 j c3k = 1 jk c11 c2 j c3k + 2 jk c12 c 2 j c3k + 3 jk c13 c 2 j c3k


= c11 (c 22 c33 c 23 c32 ) + c12 (c 23 c31 c 21c33 ) + c13 (c 21c32 c 22 c31 )
= c11

c 22

c 23

c32

c33

c12

c 21

c 23

c31

c33

+ c13

c 21 c 22
c31 c32

that is
= det(cij ) .

Exercise 19.

Given x = x i e i and y = y j e j of a Euclidean vector space, show that the expression


g ij x i y j
x i xi
is a tensor of type

y jyj

( ) and give its interpretation.


0
0

243

Tensors

Answer. From
g ij x i y j = x j y j
and
x p y p = ip jp xi y j = ij xi y j = x j y j ,
we deduce that g ij x i y j is an intrinsic scalar that is x . y ; and so are

x i xi =

x. x = x

and
y jyj = y .
The introduced expression is the cosine of angle ( x , y ) .
Exercise 20.

Given two vectors y = y i e i and z = z j e j of a Euclidean space E, we consider the


tensor of type

( ):
2
0

y z = y z z y = ( y i z j y j z i ) e i e j = t ij e i e j
that is
yz =
i< j

yi

yj

e i e j = t (ij ) e i e j .

(i) By introducing the tensor of components t ij , give an interpretation of


to the parallelogram of sides corresponding to vectors y and z.

1
t t ij
2 ij

with respect

(ii) Make explicit the contracted product t im x m between y z and any x E .


Answer. (i) We know that
t ij = g ip g jq t pq = g ip g jq ( y p z q y q z p ) = yi z j y j z i
and, of course, we have the double contracted product:
1
t t ij
2 ij

= 12 ( yi z j y j z i )( y i z j y j z i )
= t (ij ) t (ij )

that is a tensor of type

( ).
0
0

If we consider an orthonormal basis, this scalar (being independent of the basis) is expressed
as
t (ij ) t (ij ) = (t ij ) 2 = ( y i z j y j z i ) 2
i< j

i< j

which is the square of the area of the parallelogram having adjacent sides corresponding to y
and z.

244

Chapter 2

(ii) For any x = x i e i , the contracted product


t im x m = ( y i z m y m z i ) x m = z m x m y i y m x m z i
= ( z. x ) y ( y. x ) z
which defines a vector.
In the usual Euclidean space, it is the well-known formula of the triple product:
x ( y z ) = ( x .z ) y ( x . y ) z .

Exercise 21.

Given two 1-forms = i i and = j j , show various expressions of .


Answer. The tensor product
= i j i
is a

( ) -tensor such that the n


0
2

components relative to ( i j ) of T20 are i j .

The exterior product is an antisymmetric tensor of T20 :


= i j i j = i j i j i j j i

= ( i j j i ) i j
= Pij i

( where Pij =

i j
i j

),

= Pij i j + Pij i j = Pij i j + Pji j i


i< j

i j

i< j

= Pij i j Pij j i =
i< j

i< j

i< j

i j

i j
i j

i j .

To sum up:
= i j i j = ( i j j i ) i

i< j

1
( i j
2!

j i ) i j = ( i j j i ) i j .

Exercise 22.
Prove that the C np products i1 ...

ip

(i1 < ... < i p ) form a basis of p .

Answer. We have previously seen that to each p-form was associated a p-form generated

by C np elements ( i1 ...

ip

) such that i1 < ... < i p .

Thus, it is sufficient to show that these C np elements are linearly independent, that is:

245

Tensors

i1 <..<i p

i1...i p i1 ...

ip

=0

i1 ...i p = 0

(i1 < ... < i p ) .

Let ( j1 < ... < j p ) be a sequence which is completed in order that ( j1 ,..., j p , j p +1 ,..., j n ) be
(1,..., n) . The order of terms will be changed if necessary.
The assumption implies

i1 <..<i p

i1...i p i1 ...

ip

j p +1

...

jn

=0.

In this sum, only


j

j1 ... p p +1 ... jn
is nonzero, the other terms having necessarily two equal factors.

( j1 < ... < j p )

The previous sum is thus reduced to only one term, namely:

j1... j p

j1

...

jp

j p +1

...

jn

( j1 < ... < j p )

which is necessarily zero.


Thus, we conclude:
0 = j1... j p (

ji

...

jn

)(e j1 ,..., e jn ) = j1... j p

( j1 < ... < j p ) .

Exercise 23.

In a classical Euclidean plane, we consider a system of elliptic coordinates


(u , u ) ( , ) which are connected to Cartesian coordinates as follows:
1

x1 = a cosh cos

x 2 = a sinh sin

( a R+ ).

(i) Show the natural frame at any point; that is { x; e1 , e 2 } or simply { x; e i }.


(ii) Express the differential vectors de1 and de 2 in function of natural basis vectors.
Answer. By denoting x j the Cartesian coordinates of x relative to a fixed frame { o; e io }, the
position vector of x is:
ox = x j e oj

= a (cosh cos e1o + sinh sin e 2o ).


So, there are two coordinate lines through x which are:
- an hyperbola corresponding to a given value of ,
- an ellipse corresponding to a given value of .
The natural basis vectors are
x
= a (sinh cos e1o + cosh sin e 2o ) ,

x
e2 =
= a ( cosh sin e1o + sinh cos e 2o ) .

e1 =

246

Chapter 2

Fig. 67
We note that the Jacobian
J=

sinh cos
cosh sin

cosh sin
= sinh 2 + sin 2
sinh cos

is different from zero, except if ( x1 , x 2 ) is ( a,0) or (a,0) .


(ii) We have:
e
e1
d + 1 d

= a (ch cos e1o + sh sin e 2o )d + a ( sh sin e1o + ch cos e 2o )d

de1 =

= a (ch cos d sh sin d ) e1o + a ( sh sin d + ch cos d ) e 2o

and in the same manner:


de 2 = a ( sh sin d + ch cos d ) e1o + a (ch cos d sh sin d ) e 2o .

But, from
a e1o = 1 ( sh cos e ch sin e ) ,
1
2
J
1
a e 2o = (ch sin e1 + sh cos e 2 )
J

we deduce:
de1 =
de 2 =

J
1

sh ch (d e1 + d e 2 ) +

sin cos (d e1 + d e 2 )

sin cos (d e1 d e 2 ) ,
1

sh ch (d e1 d e 2 ) .

247

Tensors
Exercise 24.
By considering a transformation of curvilinear coordinates
R n R n : (u i )  (u j )

and the following tensors

a = a i jk e i e j e k ,

b = b l m e l e m ,

calculate
a:b.

Answer. The change of components of a b is such that:

a p qr b s t =

u p u j u k u s u m i
a jk b l m .
u i u q u r u l u t

A first contraction in r and s leads to

a p qr b r t =

u p u j u m k i
l a jk b l m .
i
q
t

u u u

A contraction in p and t leads to

a p qr b r p =

u j m k i l
u j i

a jk b k i .
a
b
=
jk
m
l
q i
q
u
u

This proves that a : b is a tensor of type

( ).
0
1

Exercise 25.
Let us consider a covector of Cartesian components
t1 = x1 x 3 ,

t 2 = 3( x 3 ) 2 ,

Express these components in function of spherical coordinates.

Answer. We know that


t k =

x i
ti
u k

with
u 1 = r , u 2 = and u 3 = such that:

x1 = u 1 sin u 2 cos u 3 ,
x 2 = u 1 sin u 2 sin u 3 ,
x 3 = u 1 cos u 2 .
So, we have:

t 3 = x1 x 2 .

248

Chapter 2
x1
x 2
x 3
t3
t2 +
t1 +
t1 =
u 1
u 1
u 1
= (sin u 2 cos u 3 ) x1 x 3 + (sin u 2 sin u 3 ) 3( x 3 ) 2 + cos u 2 x1 x 2
= r 2 sin 2 cos 2 cos + 3r 2 sin sin cos 2 + r 2 cos sin 2 cos sin

and so on.

Exercise 26.
Find the Christoffel symbols for spherical coordinates,
(i) from the second Christoffel formula,
(ii) from the expressions de i .

Answer. (i) From the expressions of natural basis vectors e1 , e 2 and e3 [see Example 2 of

Section 5.1.3], with spherical coordinates u 1 = r , u 2 = and u 3 = , we deduce:


g11 = e1 .e1 = 1 ,

g 22 = e 2 .e 2 = r 2 ,

g 33 = e 3 .e 3 = r 2 sin 2

and
i j : g ij = 0 .

So, we have:
ds 2 = (dx1 ) 2 + (dx 2 ) 2 + (dx 3 ) 2 = dr 2 + r 2 d 2 + r 2 sin 2 d 2

and
1

det g ij = 0

r2

( )

= r 4 sin 2 .

0 r 2 sin 2

The conjugate tensor is

0
1
r2
0

0
.

2
2
r sin

We have successively:
1
22

= g i 22 = g 122 =
1i

11

g 11
2

( 2 g12 + 2 g12 1 g 22 ) = r ,

2
2
12
= 21
= g 2i i 21 = g 22 221 =

g 22
2

( 2 g 21 + 1 g 22 2 g 21 ) =

1
,
r

Tensors

1
33

= g i 33 = g 133 =
1i

11

g 11
2

( 3 g13 + 3 g13 1 g 33 ) = r sin 2 ,

2
33
= sin cos ,
3
3
13
= 31
=

1
,
r

3
3
23
= 32
= cot .

(ii) Without introducing the fundamental tensor, from the following expressions
de1 = (cos cos e1o + cos sin e 2o sin e 3o ) d
+ ( sin sin e1o + sin cos e 2o ) d ,

de 2 = (cos cos e1o + cos sin e 2o sin e 3o ) dr


+ r ( sin cos e1o sin sin e 2o cos e 3o ) d
+ r ( cos sin e1o + cos cos e 2o ) d ,
de 3 = ( sin sin e1o + sin cos e 2o ) dr
+ r ( cos sin e1o + cos cos e 2o ) d
+ r ( sin cos e1o sin sin e 2o ) d ,

we deduce:
de1 =

d
d
e2 +
e3 ,
r
r

de 2 =

dr
e 2 r d e1 + cot d e 3 ,
r

de 3 =

dr
e 3 + cot d e 3 r sin 2 d e1 sin cos d e 2 .
r

But we know that

de j = ijk du k e i
and, by identification, we obtain for instance:
2
3
de1 = 12
du 2 e 2 + 13
du 3 e 3

and thus
2
12
=

1
,
r

3
13
=

1
,
r

1
1
1
2
2
3
3
11
= 12
= 13
= 11
= 13
= 11
= 12
=0

and so on.

249

250

Chapter 2

Exercise 27.

From the first Christoffel formula prove that

(i)

iji =

1
2

ijj =

j ln g ij ,

1
i g jj
2 g ii

in the case of systems of orthogonal coordinates.


3
2
(ii) In particular, find again 23
and 33
in spherical coordinates.

Answer. (i) We have


iji = g ih hij = g ii iij =

1 1

g ii

ijj = g ih hjj = g ii ijj =

1
g = ln g ,
ii
2 j ii 2 j

1 1

g ii 2

( i g jj ) .

(ii) We have:
3
3
23
= 32
=

2
33
=

1
2

2 ln g 33 =

1
2

ln(r 2 sin 2 ) = cot ,

1
1
2 g 33 = 2 (r 2 sin 2 ) = sin cos .
2 g 22
2r

Exercise 28.

A particle follows a circle of radius R with the constant angular speed  = 0 . From
the notion of absolute derivative, find the particle acceleration.

Answer. By considering the polar coordinates u 1 = r , u 2 = , the only nonzero Christoffel


symbols are
1
2
2
1
22
= R ,
21
= 12
=
R
The acceleration components

ai =

j
k
d 2u i
i du du
+

jk
dt dt
dt 2

are, for u 1 = r :
1
a 1 = 22

du 2 du 2
= R  2 = R 02
dt dt

2
a 2 = 21

du 2 du 1
= 0.
dt dt

and for u 2 = :

Thus, the acceleration is directed towards the circle center and its norm is proportional to the
distance from the center.

251

Tensors
Exercise 29.

Write the first Frenet formula and the components of tangential and normal
accelerations in Euclidean space.
Answer. Let =

dx
be the unit velocity where s is the curvilinear parameter of a given
ds

curve.
The corresponding components are
du i
=
ds
i

i = 1,2,3.

From
ds 2 = g ij du i du j

we deduce
g ij

du i du j
= g ij i j = 1 .
ds ds

From this result that is . = 1 , we deduce d is perpendicular to .


So, by denoting the unit vector normal to , we have the well-known formula:
d
1
=
ds R

where

is the curvature.

Explicitly, we have:
d i 1 i
= .
ds
R

The time derivative of any velocity component v i = v i =

ds i
is
dt

d i
dv
dv i
d
= (v i ) = i + v
dt
dt
dt
dt
i
dv i d 2
+
=
v
dt
ds

and thus the first formula of Frenet implies:


ai =

dv i v 2 i
+ .
dt
R

In conclusion, the acceleration a of a point following a curve shows a tangential acceleration


dv
v2
of norm
and a normal acceleration of norm
.
dt
R

252

Chapter 2

Exercise 30.

Prove that any geodesic (in a surface of classical Euclidean space) is such that at each
of its points, the osculating plane is orthogonal to the surface.
Answer. We know that every osculating plane equation is determined by the first and second
derivative vectors, and geodesic equations are
g ij

d 2u j
du k du j 1
du j du k
+

g
=0
k ij
i jk
ds ds 2
ds ds
ds 2

ei , e j

d 2u j
du j du k
1
+ ( k e i , e j i e j , e k )
= 0.
2
ds ds
ds 2

But we have:
1

k ei , e j i e j , e k = ei , k e j + e j , k ei
= ei , k e j +

[because k e i =

1
2

e j , i ek

e j , i ek
1
2

1
2

ek , i e j

ek , i e j

2 x
2 x
=
= i ek ]
u k u i u i u k

and thus the geodesic equations are written:

ei , e j

d 2u j
du j du k
+
e
,

e
=0
i
k j
ds ds
ds 2

e i , (e j

d 2u j
du j du k
+

e
) = 0.
k j
ds ds
ds 2

(1)

Let us consider the following vectors


dx du j
=
ej
ds
ds

and thus
d 2 x d 2u j
du j du k
=
e
+
ke j .
j
ds ds
ds 2
ds 2

The geodesic equations (1) mean that, at each point of the geodesic, the second derivative
d2x
is perpendicular to vectors ei .
vector
ds 2
The osculating plane is thus orthogonal to the surface.

Tensors
Exercise 31.

Let v be a vector field on a point space,


(i) Prove that the expressions

(dv ) i = dv i + v k ki
are the components of a tensor of type

( ) (called the absolute differential of v).


1
0

(ii) By considering a system of curvilinear coordinates (u j ) , prove that the expressions


j v i = v i , j + ijk v k
are the components of a tensor of type

( ) (called the covariant derivative of v ).


1
1

Answer. (i) Given a change of basis e j = ij e i , we consider


v = v i e i = v p e p .
The components of the absolute differential with respect to (e p ) are
(dv ) q = dv q + v p pq .
In order to express pq in function of ki , we consider
de p = d ( ip e i ) = d ip e i + ip ij e j = (d ip + kp ki ) e i
= (d ip + kp ki ) iq e q
but
de p = pq e q
and thus

pq = iq (d ip + kp ki ) .
So, we have:
(dv ) q = d ( iq v i ) + np v n iq (d ip + kp ki ) .
From

ip iq = pq

iq d ip = ip d iq ,

we deduce:
(dv ) q = d ( iq v i ) + np v n kp ( iq ki d kq )
= d ( iq v i ) + v k ( iq ki d kq )
= iq dv i + v i d iq v k d kq + v k iq ki
= iq (dv i + v k ki )
= iq (dv ) i .

253

254

Chapter 2

(ii) In
p vq =

v q
+ prq v r
p
u

let us make explicit prq by considering the system of curvilinear coordinates.


From (i), we know that

pq =

u q
u i
u k i
u k u q i
u q
+
(d
k ) =
(
k d k )
u i
u p u p
u p u i
u

and thus
u k u q i
2uq
(

) du j

jk
p
i
j
k
u u
u u
k
j
q
2 q
u u u i
u
prq =
( i jk j k ) .
p
r
u u u
u u

prq du r = pq =

So, we have

u k u q i
u j u q i
2 u q u r n
(
)
(
)
+

[
v

v ]
jk
u p u k u i
u r u i
u j u k u n
u k 2 u q i u q v i
u q i
2u q
(
)v j ]
=
+
+

[
v

jk
p
k
i
i
k
i
k
j
u u u
u u
u
u u
k
q
i
u u v
( k + ijk v j )
=
p
i

u u u
u k u q
=
k vi
p
i

u u
and they are really the components of a tensor of type 11 .
p vq =

()

Exercise 32.
Calculate the Laplacian of a real-valued function f in spherical coordinates u 1 = r ,

u 2 = and u 3 = .
Answer. Since
g 11 = 1 ,

g 22 =

1
,
r2

g 33 =

1
,
r sin 2
2

det g = r 4 sin 2 ,

Eq. (2-138) is written as it follows:


f =

[ 1 ( g 11 1 f r 2 sin ) + 2 ( g 22 2 f r 2 sin ) + 3 ( g 33 3 f r 2 sin )]

r sin
1
1
[ r ( r f r 2 sin ) + ( f sin ) + ( f
)]
= 2
sin
r sin
1
1 2 f
1
2 f

f
(sin
)+ 2
(r
)+ 2
= 2

r
r sin
r sin 2 2
r r
=

1 2 f cot f
1
2 f
2 f 2 f

+
+
+
+
r 2 r 2 sin 2 2
r 2 r r r 2 2

255

Tensors
Exercise 33.
Given a vector field v, find the expressions of div v :

(i) with respect to the natural basis (e r , e , e ) associated with the spherical coordinates,
(ii) with respect to the corresponding orthonormal basis (1r ,1 ,1 ) .
Answer. (i) Given
v = v r e r + v e + v e ,
we know that

vi
i (r 2 sin )
div v = i v + 2
2
r sin
i

= r v r + v + v + v r + v cot .
r

(ii) We know that


1

1r = e r ,

1 =

v r = g rr v r = v r ,

v = g v =

1 =

e ,

r sin

and
1

v ,

v = g v =

1
v
r sin 2
2

Thus, we have:
v = v r 1r +

v 1 +

r sin

v 1 .

Since the corresponding components of contravariance and covariance are identical with
respect to this (orthonormal) basis (1r ,1 ,1 ) , we deduce they are:
v = v =

vr = v r = vr ,

v ,

v = v =

1
v
r sin

and so, the well-known corresponding expression of the divergence of v is


div v = r v r +

v +

1
2
cot
v + v r +
v
r sin
r
r

Exercise 34.
Find the gradient of a function f : R 2 R in the case of elliptic coordinates u 1 = ,

u 2 = and deduce the Laplacian f .


Answer. We know that the natural basis vectors associated with elliptic coordinates and
are
e1 = a ( sh cos e1o + ch sin e 2o ) ,
e 2 = a (ch sin e1o + sh cos e 2o )

256

Chapter 2

and we have:

(g ij ) = a

( sh 2 + sin 2

,
a 2 ( sh 2 + sin 2 )
0

det g = a 4 ( sh 2 + sin 2 ) 2 ,

(g )
ij

0
2 2

a ( sh + sin )
=
.
1

0
2
2
2
+
a
(
sh

sin

From
1

( grad f ) i =

a ( sh + sin )
2

( grad f ) i =

1
a ( sh + sin 2 )
2

i f

we deduce:

det g ( grad f ) i = i f
and thus
1

f =
=

i ( det g ( grad f ) i ) =

det g

1
a ( sh + sin )
2

ii2 f

1
2 f 2 f
(
).
+
a 2 ( sh 2 + sin 2 ) 2 2

Exercise 35.

Given any two covectors x = xi e i and y = y j e j in a 3-dimensional pre-Euclidean


space, prove the formula
div( x y ) = curl x . y x . curl y ,

where we consider the adjoints of x y, curl x and curl y.


Answer. The adjoint of the
components

( ) -tensor
0
2

( x y) =
k

and thus we have:

1
2!

x y of components xi y j x j y i is the vector of

( xi y j x j y i ) =
ijk

(ij ) k
det g

( xi y j x j y i )

ijk

det g

( xi y j x j y i )

257

Tensors
div( x y ) =
=
=

det g
1

det g
1
det g

k ( (ij ) k ( xi y j x j yi ))
[ 1 ( x 2 y3 x3 y 2 ) 2 ( x1 y3 x3 y1 ) + 3 ( x1 y 2 x 2 y1 )]
[ y1 ( 2 x3 3 x 2 ) + y 2 ( 3 x1 1 x3 ) + y3 ( 1 x 2 2 x1 )
x1 ( 2 y3 3 y 2 ) x 2 ( 3 y1 1 y 3 ) x3 ( 1 y 2 2 y1 )].

This is really equal to


curl x . y x . curl y

since the following

(curl x ) k =

1
2

ijk ( i x j j xi ) =

(ij ) k
det g

( i x j j xi )

implies
curl x . y = y k (curl x ) k =

(ij ) k
det g

y k ( i x j j x i )

and in the same manner:


x . curl y =

(ij ) k
det g

x k ( i y j j y i ) .

Exercise 36.

Calculate the components of the gradient of a real-valued function f and the


corresponding contravariant components (under g# )
(i) relative to the natural basis (e r , e , e z ) of cylindrical coordinates,
(ii) relative to the basis of unit vectors (1r ,1 ,1z ).
Answer. (i) With respect to the natural basis (e r , e , e z ) of cylindrical coordinates, the
components of covector grad f are
r f , f , z f

and since g 11 = g 33 = 1 and g 22 = 1 r 2 , we obtain:


( grad f )1 = g 11 1 f = r f ,
( grad f ) 2 = g 22 2 f =

f ,
r2
( grad f )3 = g 33 3 f = z f .

258

Chapter 2

(ii) We know that the natural basis (e r , e , e z ) with u 1 = r , u 2 = , u 3 = z is such that


e r = cos e1o + sin e 2o ,

e = r sin e1o + r cos e 2o ,

e z = e 3o

thus

1r = e r ,

1 =

1z = e z .

e ,

The components ( grad f ) i relative to (1r ,1 ,1z ) are respectively:


r f ,

1
f , z f .
r

Indeed, the following point cannot be too strongly emphasized: We know


df = ( grad f ) . dx ,

and with respect to the natural basis, we have:


dx = dr e r + d e + dz e z
and
df = i f du i = r f dr + f d + z f dz .
With respect to (1r ,1 ,1z ) , we have:
dx = dr 1r + r d 1 + dz 1z
and thus the corresponding components ( grad f ) i are respectively:
r f ,

1
f , z f .
r

They are the components (grad f ) i too, since (1r ,1 ,1z ) is an orthonormal basis.
Exercise 37.

Given the cylindrical coordinates r , , z :


(i) Find the only nonzero Christoffel symbols with respect to the corresponding (orthonormal)
basis (1r ,1 ,1z ) also denoted (e r , e , e z ) :
r =

r =

and prove, for example:


r = 0.

(ii) Given (e r , e , e z ), calculate the divergence of a tensor t of type

( ).
2
0

Answer. (i) We are going to calculate the previous Christoffel symbols with respect to
(1r ,1 ,1z ) from the known Christoffel symbols with respect to the natural basis (e r , e , e z ).

259

Tensors

Since in general
ijk = k e j , e i ,

we have the following results for the particular basis (e r , e , e z ) [see Sect. 5.6.1a]:
e , (e )
r r

r =

r =

= e r , d = e r , d = r = 1 r ,

e , (e ) r =

r
( e ), dr = 2
= ,
r
r
r
r

r = r e , (e ) = r ( e ), r d = 2 e + r e , r d
r
r
r
=

e , d + r e , d = + r = 0 .
r
r

(ii) From expression (2-133) of the divergence, we obtain the following components relative
to (e r , e , e z ) :

for i = r :
1

(div t ) r = r t rr + t r + z t rz + r t + r t rr
r
1
t rr t
rr
r
zz
= r t + t + z t +
,
r
r
for i = :
1

(div t ) = r t r + t + z t z + r t r + r t r
r
1

= r t r + t + z t z + t r ,
r
r

for i = z :
1

(div t ) z = r t zr + t z + z t zz + r t zr
r
1

= r t zr + t z + z t zz + t zr .
r
r

Exercise 38.

Find the geodesic equations of a torus T 2 in the usual Euclidean space E.


Answer.
Let ox1 x 2 x 3 be a fixed system of reference of E,
c(o, R1 ) be a horizontal circle of center o and radius R1 ,
c(O, R2 ) be a circle of center O c(o, R1 ) situated in a vertical plane through o.

260

Chapter 2

Fig. 68

Let Ox 1 x 2 x 3 be the system of reference ensuing from a rotation of angle about axis ox 3 ,
where is the angle between the axes ox1 and ox 1 , the axis ox 3 being parallel to ox 3 .
The coordinates of any point of the vertical circle c(O, R2 ) in ox 1 x 2 x 3 are
x 1 = R1 + R2 cos ,

x 2 = 0 ,

x 3 = R2 sin

where designates the angle locating the point on the circle c(O, R2 ) .
The coordinates of this point with respect to the fixed system of reference are
x1 cos

x 2 = sin

x3 0

sin
cos
0

1
0 x

0 x 2 ,

1 x 3

which implies the following parametric equations of T 2 :


x1 = ( R1 + R2 cos ) cos ,

x 2 = ( R1 + R2 cos ) sin ,

x 3 = R2 sin

where the longitude [0 , 2 ) and latitude [ 0, 2 ) are the coordinates on T 2 .


Since
ds 2 = ( R1 + R2 cos ) 2 d 2 + ( R2 ) 2 d 2 ,

the Christoffel formula implies:


=
= 0,


=
= 0,

( R1 + R2 cos ) sin
,
R2

R2 sin

R1 + R2 cos

261

Tensors

The geodesic equations are

&& +

R2

&&

( R1 + R2 cos ) sin & 2 = 0 ,

2 R2 sin & &


= 0 ,
R1 + R2 cos

( R2 ) 2 & 2 + ( R1 + R2 cos ) 2 & 2 = 1 .

These equations are not independent. The second equation leads to a first integral
( R1 + R2 cos ) 2 & = k

(k R) .

Exercise 39.

Given the spherical coordinates r , , and the corresponding orthonormal basis


(e r , e , e ) :
(i) Prove that the only nonzero Christoffel symbols are
r = r = r = r =
= =

cot
.
r

1
r

(ii) Given (e r , e , e ) calculate the divergence of a tensor of type

( ).
2
0

Answer. (i) Since in general


ijk = k e j , e i ,
we have [see Sect. 5.6.1a] the following results:
1

r =

e r , (e ) =

r sin

1
r

1
,
r

1
1
( e ), r sin d
r sin
r
cot
=
r

e , (e ) =

1
e , d =
r
r

e r , r d = e r , d = r =

and so on.
(ii) From expression (2-133) of the divergence, we obtain the following components with
respect to ( e r , e , e ) :
for i = r :
1
1
t r + r t + r t + r t rr + r t rr + t r
(div t ) r = r t rr + t r +
r
r sin
1
1
1
= r t rr + t r +
t r + (2t rr t t + t r cot ) ;
r
r sin
r

262

Chapter 2

for i = :
1
1
(div t ) = r t r + t +
t + r t r + t + r t r + r t r + t
r
r sin
1
1
1
= r t r + t +
t + (3t r + (t t ) cot ) ;
r
r sin
r

for i = :
1
1
(div t ) = r t r + t +
t + r t r + t + r t r + r t r + t
r
r sin
1
1
1
= r t r + t +
t + (3t r + 2t cot ) .
r
r sin
r

CHAPTER 3

MASS GEOMETRY
INERTIA TENSOR

The mass geometry as well as the inertia tensor are introduced in the first course of
mechanics at the undergraduate level; but at this level, the inertia tensor is often studied in the
matrix context and it is the reason why we introduce this notion in the frame of the tensor
theory.
In mechanics, the inertia tensor is the first tensor of second order met by students in all
probability. It is an essential element for the study of dynamics systems. In addition, we recall
this important notion because the methods used to find principal axes, inertia ellipsoids, etc.
are also encountered in continuum mechanics by considering the stress tensor.

1. MASS DISTRIBUTION AND INTEGRALS


Previous notations of spaces are used again.
Let us consider a set of particles which constitute a material system S.
1.1

DENSITY

Let D denote the set of points locating the particles in the frame of reference and
simply called domain of S.
We know that a material system S can be considered as discrete or continuous. We recall the
following notions.
Given a discrete system S of N particles, that consists in N points x1 ,..., x N with respective
masses m1 ,..., m N , we say:

263

264

Chapter 3
The mass of S is the real
N

m(S ) = mh

( 0 ).

(3-1)

h =1

Given a continuous system S, if D stands for the 3-dimensional domain of any subsystem S
of S ( D D ), then we say:
D

The mass of S is the triple integral

m(S ) =

D'

dV

(3-2)

where the (piecewise) continuous function

: D R+ : x  ( x )
is called the density or mass per unit volume (or specific mass).
Given a point x belonging to some set D ( D) of volume V and of mass m , we define:
D

The density at x is the limit1


lim

V 0

m d m
=
= ( x) .
V dV

(3-3)

The density at every point shows the mass distribution.


The dimension of is [ ML3 ] , (kg m 3 ).
Any real material system occupies a 3-dimensional domain, but the following cases of
approximation can be considered as above:
(i) If S is the 2-dimensional domain of any subsystem
double integral

of S, then the mass of

is the

m(S ' ) = dA
S

where

: S' R+ : x  ( x)
is the mass per unit area.
Given a point x belonging to some set S ( S ) of area A and of mass m , we define the
density at x as the limit
lim

A0

m d m
=
= ( x) .
A dA

More precisely, in this definition we assume that the diameter of

D tends to zero.

Mass Geometry, Inertia Tensor

(ii) If C is the one-dimensional domain of any subsystem


curvilinear integral

265

S of S, then the mass of S is the

m(S ) = ds
C

where

: C R+ : x  ( x )
is the mass per unit length.
Given a point x belonging to some set C ( C ) of length s and of mass m , we define
the density at x as the limit
lim

s 0

m d m
=
= ( x) .
s
ds

General notation. By considering


dm = ( x) d
N

and by interpreting m = mi as a positive measure, all the previous integrals can be


i =1

described as the following integral:

dm

(3-4)

viewed as a Stieltjes integral1 and where we identify the integration domain with the material
system in order to simplify the notation.

1.2

INTEGRALS OF REAL-VALUED AND VECTOR FUNCTIONS

First, let f : D R : x  f ( x) be a (piecewise) continuous function and be a


density in the 3-dimensional domain D .
We assume that triple integrals of type

f ( x) ( x) dV

exist.
In particular, there are double integrals of type

f ( x) ( x) dA

or curvilinear integrals of type

C
1

f ( x) ( x) ds .

For a continuous mass distribution, it is a Riemann integral, while for a system made up of a finite number of
particles the integral is replaced by a sum. The Stieltjes integral lets a unique presentation of various cases.

266

Chapter 3

Given a material system S, the domain of which is D, S or C, then the previous integrals let
define:
D  The integral of f with respect to a mass distribution of S is

f dm

where dm = dV (or dA , or ds ).
The following properties are fulfilled:
(i) Additivity : Given domains Dk (k = 1,..., p) of various subsystems S k of a material
system S such that
p

m(S ) = m(S k ) ,
k =1

then we have:

f dm =

k =1

f dm .
k

(ii) Linearity: Given two functions f, g and two reals a, b, we have:

S
(iii) Positivity :

(af + bg ) dm = a f dm + b g dm .
S

[ x : f ( x) 0 ]

f dm 0 .

We also recall that the mass of S is obtained for f = 1 :


m(S ) = dm .

(3-5)

Now, we define a vector function on a 3-dimensional domain D


f : D E : x  f ( x)

from the three components f1 ( x), f 2 ( x) and f 3 ( x) of the vector f (x) with respect to a basis
(e1 , e 2 , e 3 ) of a 3-dimensional vector space E.

We set:
D

The integral of a vector function f over S is such that

f ( x) dm =
i =1

f i ( x) dm e i .

From the linearity of real-valued functions, the reader will verify that this definition is
independent of the choice of basis.
The properties of additivity and linearity are immediate.

267

Mass Geometry, Inertia Tensor

2. CENTER OF MASS
2.1

DEFINITIONS

First of all, let us consider a system S of N particles p h such that each particle is
defined by a point xh with a positive real mh (the mass of p h ).
The mass of the material system S is denoted
N

M = mh .
h =1

The center of mass of S, say G, is defined by the position vector


oG =

1
M

mh ox h .

(3-6)

h =1

Remark 1. The previous definition does not depend on the choice of the reference point o.

Indeed, from
o1 o + oG1 = o1G1 =

1
M

h =1

mh o1 x h =

1
M

mh (o1 o + ox h )
h =1

and since
o1 o =

1
M

mh o1 o ,
h =1

we deduce:
oG1 =

1
M

mh ox h
h =1

which implies that G1 and G coincide.


Remark 2. We recall that in a uniform gravitational field the center of mass is the center of
gravity.
An intrinsic definition of G is the following.
D

The center of mass of S is defined independently of any reference point as follows:


N

mh Gx h = 0 .
h =1

Both the definitions are equivalent because we have successively:

(3-7)

268

Chapter 3
N

mh Gx h = 0
h =1

h =1

h =1

Go mh + mh ox h = 0
N

M oG = mh ox h .
h =1

Now, let S be a continuous system made up of particles.


We denote by r = op the position vector of any particle.
D

The center of mass of S, say G, is defined by the position vector


oG =

1
M

r d

(3-8)

where D denotes a 3-dimensional domain of S.

In particular, if S is a 2-dimensional domain of a continuous material system, we define G as


follows:
oG =

1
M

S r dA .

If C is a one-dimensional domain of a continuous material system, we define G as follows:


oG =

1
M

C r ds .

We sum up all the previous definitions of G:


D


The center of mass G of a material system S is defined by the position vector


oG =

M S

r dm

(3-9)

where r ox locates every particle of S,


or intrinsically:


S Gx dm = 0 .

The two previous expressions of the center of mass are equivalent since

S Gx dm =S Go dm + S r dm = M oG + S r dm = 0 .

(3-10)

269

Mass Geometry, Inertia Tensor


Coordinates of G.
Given a coordinate system of reference, the coordinates of G are
xG =

x dm
M S

yG =

M S

y dm

zG =

M S

z dm

where dm = d (that is dV , dA or ds according to the case). We have in particular:


xG =

mh x h
M

yG =

h =1

mh y h
M
h =1

zG =

mh z h .
M
h =1

Example. Determine the center of mass of a thin circular plate S of radius R and of density

= kr n (2 cos )
where r and are the polar coordinates.
Answer. The x-axis is chosen as the axis of symmetry since (r , ) = (r , ) and thus
yG = 0 .
The other coordinate of the mass center is
xG =

M S

x dm

with
dm = dA = kr n (2 cos )r dr d .

The mass of the plate is


R

M = k r n+1dr

(2 cos ) d =

k 4 R n + 2
n+2

and since
xG =

M S

r cos dm =

k R n +3
,
M n+3

we have:
xG =

2.2

1 n+2
4 n+3

R.

SUBDIVISION
Let S be a material system formed by p subsystems S1 ,..., S p with respective centers

of mass G1 ,..., G p and with masses M 1 ,..., M p respectively.


PR1

The center of mass of S is the center of mass of the system made up of centers of mass
of different subsystems with their respective masses:

270

Chapter 3

oG =

1
M

M i oGi

(3-11)

i =1

Proof. We have clearly:


n1

mh ox h = M 1 oG1 ,

ox d = M 1 oG1 ,

ox d = M 2 oG 2 ,

h =1

n2

mh ox h = M 2 oG 2 ,

h = n1 +1

.
.

.
.

np

mh ox h = M p oG p .

h = n p 1 +1

ox d = M p oG p .

The sum of all the left-hand members being


np

mh ox h = M oG ,

D ox d = M oG ,

h =1

we conclude that
M oG =

M i oGi .
i =1

Example. Find the center of mass of the homogeneous plate illustrated in Fig. 69.

Fig. 69
We subdivide the plate into three rectangular parts. For the homogeneous plate, Eq. (3-11) is
written:

oG =

Ai oGi
i =1

where the various Ai are the areas of the respective parts.

271

Mass Geometry, Inertia Tensor

By choosing the axis of symmetry as x-axis and the y-axis as in the figure, we obtain
immediately:
xG =

A1 xG1 + A2 xG2 + A3 xG3


A1 + A2 + A3

yG = 0

with

xG1 = 2.510 2 (m),

xG3 = 0.21 (m),

xG2 = 0.1 (m),

and thus

xG = 3.5625 10 2 (m).

2.3

THEOREMS OF GULDIN (PAPPUS)

Let C be a homogeneous curve of length l revolving about a nonintersecting x-axis


situated in the (x,y)-plane of the curve.
PR2

A plane curve which rotates about an axis in its own plane (and does not intersect the
axis) generates a surface whose area A is equal to the length of the curve multiplied by
the length of the circle described by the center of mass of the curve.

Proof. By denoting yG the ordinate of the center of mass of C and M being the mass of C,
the generated lateral area is
A = 2 y ds = l 2
C

M C

y ds

= l 2 yG .

Now we consider a plane surface S under the curve C and of area A.


PR3

A plane surface under a curve, by revolving about an axis (which does not intersect the
curve), generates a volume which is equal to the area of the surface multiplied by the
length of the circle described by the center of mass of the surface.

Proof. By denoting yG the ordinate of the center of mass of the plane surface S and M
being the mass of S, the volume generated by the rotation is
b

V = y 2 dx = A
a

= A 2 y G .

(
A a 0

2 y dy ) dx = A

M S

M
A

y dx dy

272

Chapter 3

3. INERTIA TENSOR
3.1

MOMENTS AND PRODUCTS OF INERTIA

With respect to a coordinate system oxyz and given a mass distribution of a material
system S, we say:
D

A moment of order k is the integral

S x

y q z r dm

where k = p + q + r .
So, the only moment of order zero is the mass

M = dm .
S

The three moments of order 1 are connected to the coordinates of the center of mass; namely:

M xG = x dm ,
S

M y G = y dm ,

M z G = z dm .

The six moments of order 2 are:


-

The moments of inertia with respect to the coordinate planes, namely:

I yoz = mh x h2 ,

(3-12)

I yoz = x 2 dm ,

I zox = mh y h2 ,

(3-13)

I zox = y 2 dm ,

I xoy = mh z h2 ,

I xoy = z 2 dm .
S

(3-14)

Pxy = mh x h y h ,

Pxy = xy dm ,
S

(3-15)

Pyz = mh y h z h ,

Pyz = yz dm ,
S

(3-16)

Pxz = mh x h z h ,

Pxz = xz dm .

(3-17)

The products of inertia, namely:

In a general way, we introduce the following important definition of moment of inertia which
characterizes the distribution of masses in the material system.

273

Mass Geometry, Inertia Tensor


D

A moment of inertia of S with respect to a point, a straight line or a plane is

mh rh2

S r

dm

(3-18)

where rh (or r) is the distance between any p h (or p) of S and the point, straight line
or plane respectively.
The moments of inertia are all the more important since the particles are farther and have
greater masses.
The dimension of any moment of inertia is [ ML2 ] .
The moments of inertia are respectively
-

about the origin:


I o = mh ( x h2 + y h2 + z h2 ) ,

I o = ( x 2 + y 2 + z 2 ) dm ,

(3-19)

I x = mh ( y h2 + z h2 ) ,

I x = ( y 2 + z 2 ) dm ,

(3-20)

I y = mh ( z h2 + x h2 ) ,

I y = ( z 2 + x 2 ) dm ,

(3-21)

I z = mh ( x h2 + y h2 ) ,

I z = ( x 2 + y 2 ) dm ,

(3-22)

- about the coordinate axes:


S

- about the coordinate planes are I yoz , I zox and I xoy .


Remark 1. There are relations between the seven previous moments as
I x = I xoy + I zox ,

I y = I yoz + I xoy ,

I z = I zox + I yoz ,

I o = I xoy + I yoz + I zox = 12 ( I x + I y + I z )


= I x + I yoz = I y + I zox = I z + I xoy .

(3-23)
(3-24)

From these relations, we deduce that

I xoy = 12 ( I x + I y I z ) .

(3-25)

Remark 2. Every product of inertia can be expressed in function of moments of inertia.

Indeed, since the squares of distances from particles to bisecting planes of respective
equations y z = 0 and y + z = 0 are
2

d h = 12 ( y h z h ) 2 ,

d h = 12 ( y h + z h ) 2 ,

we deduce for instance:


Pyz = mh y h z h = 12 [ mh d h 2 mh d h 2 ] .
h

274

3.2

Chapter 3
INERTIA TENSOR

We are going to introduce a tensor the components of which are the moments of
inertia of S about the coordinate axes and the products of inertia with the minus sign.
Given a system of coordinates ox1 x 2 x 3 in the usual 3-dimensional Euclidean space, we say:
D  The tensor of inertia of S is a symmetric tensor of second order such that

I ij = ( g ij g rs g ir g js ) x r x s dm .

(3-26)

It is denoted by I.
In particular, for a discrete system of N particles, we have:
N

I ij = mh ( g ij g rs g ir g js ) x(rh ) x(sh ) .

(3-27)

h =1

We note that the components of the inertia tensor are obviously written:
I ij = ( g ij x s x s xi x j ) dm .

(3-26)

Given an orthonormal basis, we deduce:




I ij = ( ij x r x r xi x j ) dm

).

(3-28)

So, we have:
I11 = ( 11 x r x r x1 x1 ) dm = (( x 2 ) 2 + ( x3 ) 2 ) dm = I x1 ,
S

I12 = ( 12 xr x r x1 x 2 ) dm = x1 x 2 dm = Px1x2
S

and so on.
In conclusion, given the previous frame such that x1 = x1 = x , x 2 = x 2 = y , x 3 = x3 = z , the
inertia tensor is defined by
I x Pxy Pxz

Pyx I y Pyz .

Pzx Pzy
I z

Tensors can be represented by matrices in any coordinate system, but it is obvious that
matrices do not fulfill the tensor rules.

275

Mass Geometry, Inertia Tensor

Given a system of (curvilinear or rectilinear) coordinates, we recall that the components of a


tensor of second order are transformed according to the following rule
I pq =

u i u j
I ij .
u p u q

(3-29)

Explicitly, this is denoted by

I12

I11

I 22

I 21
I I
31 32

u 1

u 1

I13
1

u
= 2
I 23
u
1
I 33
u
u 3

u
u

u
u

3
1
3
2

3
3

u 1

u 1
I
I
I

11 12 13 2
I 21 I 22 I 23 u
1

u
I
I
I
31 32
33
3

u 1

1
u

u
u

u
u

3
u
3
u
3
u
u

or briefly:
t

u
u
I =
.
I
u
u
'

(3-29)

Example. We consider a rotation of angle about the z-axis. If we let x1 = x, x 2 = y and

x 3 = z , the rotation is expressed as follows:


x1 cos sin

x 2 = sin cos

0
x3 0

1
0 x

0 x 2 .

1 x 3

We obtain:
x i
x 1
x1
=
x 1

x j
I ij
x 1
x1
x1 x 2
x 2 x1
x 2 x 2
I
+
I
+
I
+
I 22
11
12
21
x 1
x 1 x 1
x 1 x 1
x 1 x 1
= cos 2 I 11 + cos sin I12 + sin cos I 21 + sin 2 I 22

=
I11

that is

I x = cos 2 I x + 2 sin cos ( Pxy ) + sin 2 I y .


In the same manner:
x i x j
I ij
x 2 x 2
x 1 x 1
x 1 x 2
x 2 x 1
x 2 x 2
I 11 +
I 12 +
I 21 +
I 22
=
x 2 x 2
x 2 x 2
x 2 x 2
x 2 x 2
= ( sin )( sin ) I x + ( sin cos )( Pxy ) + cos ( sin )( Pxy ) + cos 2 I y

=
I 22

276

Chapter 3

that is

I y = sin 2 I x 2 sin cos ( Pxy ) + cos 2 I y .


In the same manner:
x i
x 1
x1
=
x 1

x j
I ij
x 2
x1
x1 x 2
x 2 x1
x 2 x 2
I
+
I
+
I
+
I 22
11
12
21
x 2
x 1 x 2
x 1 x 2
x 1 x 2
= cos ( sin ) I11 + cos 2 I12 + sin ( sin ) I 21 + sin cos I 22

=
I12

that is
Pxy = sin cos ( I y I x ) + (cos 2 sin 2 )( Pxy ) .
We note that the moments and products of inertia for which the coordinate x 3 is present are
unchanged since the distances with respect to the z-axis and to the plane xoy are not
modified.
We also mention that the last equality lets find new axes for which the product of inertia Pxy
vanishes since

sin cos
cos 2 sin 2

Pxy
I y Ix

which amounts to a rotation about the z-axis such that

= 12 tan 1

2 Pxy
Iy Ix

4. INERTIA ELLIPSOID
Given an orthonormal basis (e1 , e 2 , e 3 ) associated with a frame oxyz , we consider a
straight line through o.
Given a material system, we are going to prove that a very important quadric can be obtained
by considering the moment of inertia about any axis through o.
4.1

cos .

MOMENT OF INERTIA ABOUT AN AXIS

The components of a unit vector e along are the direction cosines cos , cos and

Let S be a system of particles which is not a straight line of mass points.

277

Mass Geometry, Inertia Tensor

PR4

The moment of inertia about an axis (in a direction defined by e) is expressed as


I = e .I e

(3-30)

which is the contraction between e and I e .

Fig. 70
Proof. Let ox h be the position vector of any particle p h .

The square of the distance of p h from is


d h 2 = ox h

ox h

where ox h is the vector projection of ox onto .


Let x h , y h , z h be the components of ox h .
From
ox h
ox h

= xh 2 + yh 2 + z h 2 ,

= (ox h . e ) 2 = ( x h cos + y h cos + z h cos ) 2 ,

we deduce that the inertia moment of S with respect to is


I = mh [ x h 2 + y h 2 + z h 2 ( x h cos + y h cos + z h cos ) 2 ]
h

and since cos 2 + cos 2 + cos 2 = 1 , we obtain a quadratic form E3 R : e a I such


that:
I = cos 2

mh ( y h 2 + z h 2 ) + cos 2 mh ( z h 2 + xh 2 ) + cos 2 mh ( xh 2 + y h 2 )
h

2 cos cos

mh y h z h 2 cos cos mh z h xh 2 cos cos mh xh y h ;


h

278

Chapter 3

that is
I = I x cos 2 + I y cos 2 + I z cos 2
2 Pyz cos cos 2 Pzx cos cos 2 Pxy cos cos .

(3-31)

cos

By considering the vector e = cos , we have:


cos

I x Pxy Pxz cos

I = (cos cos cos ) Pyx I y Pyz cos

Pzx Pzy I z cos

which is really
I = e .I e .

4.2

EQUATION OF THE QUADRIC

In order to establish the equation of the inertia ellipsoid of S with respect to o, we


consider, for any , the point p of position vector:

op =

e
I

( I > 0 ).

The coordinates ( x, y, z ) of this point p [with respect to the orthonormal basis (e1 , e 2 , e 3 ) ] are
written:
cos
cos
cos
x = op . e1 =
,
y=
,
z=
.
I
I
I
So the expression (3-31) of I leads to the equation of a quadric of center o:
I x x 2 + I y y 2 + I z z 2 2 Pyz yz 2 Pzx zx 2 Pxy xy = 1

(3-32)

that is, by letting x1 = x , x 2 = y , x 3 = z :




I ij x i x j = 1 .

Remark 1. The reader will immediately bring together

I = e .I e
and
I = I ij ( I x i )( I x j )
where the expressions between brackets are the components of e.

(3-32)

Mass Geometry, Inertia Tensor

279

Remark 2. We note that there was no dimensional homogeneity when we defined op leading
to the equation of the quadric. In fact, we must introduce a scale factor a for dimensional
reasons:
op =

ae
I

But it is usual to set this factor equal to 1. So, the construction of the ellipsoid, but not its
shape is subject to the choice of units.
Remark 3. We emphasize that in every change of coordinates the coefficients of a quadric
( I ij x i x j = 1 here) are the covariant components of a symmetric tensor of second order (the
inertia tensor here).

Indeed, first given any change of coordinates


x i = ki x k ,

the equation of the quadric becomes


I ij ki rj x k x r = 1 .

Since
x k x r = 1 ,
I kr
we deduce that
= ki rj I ij .
I kr
Inversely, the various components of a symmetric tensor ( I ij of the inertia tensor here) define
the coefficients of a quadric (of equation I ij x i x j = 1 here).

4.3

NATURE OF THE QUADRIC

We are going to specify the nature of the quadric of center o which is associated with
the material system S whatever the system of coordinates.
In a general manner, the quadric is an ellipsoid.
Indeed, if all the particles do not belong to a straight line through o, then op = 1 I is
always finite and we conclude that the quadric is an ellipsoid since it is the only quadric
having no point to infinity and this quadric is called: the inertia ellipsoid.
In the particular case where all the mass of a material system is distributed along a straight
line through o, for example the z-axis, then the only nonzero components of the inertia tensor
are
I x = I y = z 2 dm .
S

280

Chapter 3

Except for z-axis, the previous construction of the quadric is possible and, since
I = I x (cos 2 + cos 2 ) ,

we obtain the following equation of the quadric:


I x (x2 + y 2 ) = 1 .

It is the equation of a cylinder of revolution about the z-axis.


In the particular case where all the mass of a material system is distributed in a plane, for
instance the (y,z)-plane, then we have:
Pyz = Pzx = 0

and
Iz = Ix + Iy

since
I x = y 2 dm ,

I y = x 2 dm .

Thus, the equation of the inertia ellipsoid is


I x x 2 + I y y 2 + ( I x + I y ) z 2 2 Pxy xy = 1 .

So, the intersection of this ellipsoid with the (y,z)-plane is the inertia ellipse of equations:
I x x 2 + I y y 2 + ( I x + I y ) z 2 2 Pxy xy = 1

z = 0 .

4.4

RADIUS OF GYRATION

The radius of gyration of a system S with respect to a straight line is the distance K
from such that
I = M K 2

K = I M .

(3-33)

In other words, K is the distance such that, if all the mass of S were situated at a distance K
from , then the moment of inertia would be I .
Example. Express the radius of gyration of a homogeneous cylinder of mass M, height h and
radius R, with respect to the axis of the cylinder (z-axis).

We have immediately:
I z = ( x 2 + y 2 ) dm =
S

and thus
K=R

2.

2 r 3 h dr =

hR 4 = M R 2 2

281

Mass Geometry, Inertia Tensor

5. PRINCIPAL AXES
We know that a tensor which is in diagonal form in a particular coordinate system will
generally no longer be so after a change of coordinates; but the following fundamental
proposition can be proved.
We are going to consider any symmetric second-order tensor T that will be the inertia tensor
in particular.
5.1

FUNDAMENTAL THEOREM ABOUT SYMMETRIC TENSORS

PR5

Every real and symmetric tensor can be brought into diagonal form by an orthogonal
transformation.
The resulting diagonal tensor is unique except for the presentation order of its diagonal
elements.

In other words, this proposition states that by a suitable choice of coordinate axes a symmetric
tensor can be made diagonal.
Diagonalization of a tensor is an eigenvalue problem of linear algebra and we recall:
D

An eigenvector of a tensor T is any vector v parallel to the vector T v , that is:


T v = v

(3-34)

where the real is the corresponding eigenvalue.


Before demonstrating the fundamental proposition, we are going to prove the following
property.
P1

The eigenvalues of a symmetric real tensor are real.

Proof. By considering T v = v , first we let the various eigenvalues and components of


corresponding v be complex.

[The complex vector norm is v =

v .v where v denotes the complex conjugate of v.]

The complex conjugate of the following equation

v .T v = v . v

(1)

v .T v = v . v

(2)

is
(because T is real and the scalar product is commutative).
Since, from the definition of the transpose t T of T , we deduce:
( v T ) . v =( t T v ) . v = v . t T v ,

and since T is symmetric, we have obtained:

282

Chapter 3

v T . v = v .T v ;

that is (1) = (2) and thus any eigenvalue is such that

= .
Now, we prove the proposition by following the general method of obtention of principal
axes; that is, by making explicit
(T I ) v = 0 ,

(3-34)

(T11 ) v1 + T12 v 2 + T13 v3 = 0 ,

T21 v1 + (T22 ) v 2 + T23 v3 = 0 ,


T v + T v + (T ) v = 0 .
33
3
31 1 32 2

(3-35)

namely:

This linear system of equations has a nontrivial solution provided:


det(T I ) = 0 .

This cubic equation called the characteristic equation of the tensor has in general three
solutions: the three eigenvalues (1) , ( 2) and (3) .
Any eigenvalue is substituted in the system of equations (3-35), which leads to the direction
of the corresponding eigenvector v (only the direction since we obtain the ratios of values
v1 , v 2 and v 3 ).
For each eigenvalue (1) , ( 2) , (3) we consider a unit vector E1 , E 2 , E 3 respectively along
the direction of the corresponding eigenvector, and we say:
D

The axes defined by the unit vectors E i along the directions of eigenvalues of T are
called the principal axes of T.

So, we emphasize that:




The tensor T is in diagonal form in the frame of its principal axes.

We must prove again that the previous vectors E j are perpendicular; that is:
P2

The eigenvectors of a symmetric tensor corresponding to different eigenvalues are


orthogonal.
In other words:
The principal axes of T are associated with the vectors E i of an orthonormal basis.

Proof. Given two eigenvalues (i ) ( j ) of T ; that is:


T v (i ) = ( i ) v (i ) ,

we have:

T v ( j ) = ( j ) v ( j ) ,

283

Mass Geometry, Inertia Tensor


v ( j ) .T v ( i ) = ( i ) v ( j ) . v ( i ) ,
v ( i ) .T v ( j ) = ( j ) v (i ) . v ( j ) .

The left-hand members of previous equations are equal since

v ( j ) .T v ( i ) = ( t T v ( j ) ) . v ( i ) = v ( i ) .T v ( j ) .
Therefore, from
( ( i ) ( j ) ) v ( i ) . v ( j ) = 0

we deduce:
v (i ) . v ( j ) = 0

since (i ) ( j ) .
To conclude, we have obtained vectors E j which are expressed with respect to the vectors of
the frame of reference as follows:

E j = ij e i
where the various ij = E j . ei are the coefficients of the orthogonal transformation from
{ o; e i } to { o; E j }.
The so-proved fundamental proposition can be particularized in the inertia tensor study; the
principal axes are called the principal axes of inertia of S (at point o).
PR6

The eigenvalues of the inertia tensor are the moments of inertia about the principal
axes of inertia.

Proof. It is obvious since if is the axis defined by a unit eigenvector E corresponding to an


eigenvalue , then we have:
E .I E = E . E = .

The expression of the inertia tensor, at o, with respect to the principal X,Y Z-axes, is
I X 0 0
0 I 0 .
Y

0 0 I Z

(3-36)

A moment of inertia about a principal axis is called a principal moment of inertia.

PR7

Any principal moment of inertia cannot be higher than the sum of the two other ones.

Proof. It is obvious from the definition of the principal moment of inertia.

284

5.2

Chapter 3
EQUAL EIGENVALUES

Until now, we have dealt with the case of different eigenvalues. The so-obtained
respective principal axes are unique except for degeneracy.
So, the cases of double and triple roots of the characteristic equation must be to consider.
The following developments are valid (as before) for any real symmetric second-order tensor;
but we will particularize this tensor by considering the inertia tensor I.
- First, we consider a double root: (i ) = ( j ) ( k ) .
For instance, the principal axis corresponding to the eigenvalue (3) is chosen as Z-axis
without restriction.
PR8

If two eigenvalues are equal ( (1) = ( 2) ), then the inertia tensor is of type

I X 0 0

0 I
X 0 .

0 0 I Z

(3-37)

Proof. In this case where v = E 3 , in the system of equations (3-35) we successively have
v1 = cos = 0, v 2 = cos = 0, v3 = cos = 1 , and this system becomes:

( I X )cos PXY cos PXZ cos = PXZ = 0 ,

PXY cos + ( I Y ) cos PYZ cos = PYZ = 0 ,


P cos P cos + ( I ) cos = I = 0.
YZ
Z
Z
XZ
So, we have obtained:

( 3) = I Z ,

PXZ = PYZ = 0 .

Conversely, if the products of inertia PXZ and PYZ are zero, then the Z-axis is a principal axis
of inertia.
So, the inertia tensor is a priori:
IX
P
XY
0

PXY
IY
0

0
0
I Z

with the corresponding characteristic equation:


2
( I Z ) [( I X )( I Y ) PXY
] = 0.

The eigenvalues (1) and ( 2) are the roots of


2
2 ( I X + I Y ) + I X I Y PXY
= 0.

Mass Geometry, Inertia Tensor

285

Since these two roots are equal, the discriminant vanishes:


2
( I X I Y ) 2 + 4 PXY
=0

which implies:
I X = IY ,

PXY = 0

and the predicted inertia tensor is found.


PR9

Given a double eigenvalue, every unit vector E perpendicular to the principal vector
corresponding to the other eigenvalue is an eigenvector associated with the double
eigenvalue.

Proof. If (1) = ( 2) , we choose the eigenvector corresponding to (3) along the Z-axis.
We have:
I X 0 0 cos

I E = 0 I X 0 cos = I X E
0 0 I 0
Z

where the expression of E with respect to an orthonormal basis shows no component along
the E3-axis.
We have proved that E is the eigenvector associated with I X .
The existence of principal axes of I follows from the preceding proposition.
-

Second, we consider a triple root: (i ) = ( j ) = ( k ) .

We have immediately:
PR10 In the case of a triple eigenvalue, every unit vector E is a unit eigenvector associated
with the triple eigenvalue.
Proof. This is obvious from the previous case where I X = I Z and the inertia tensor is
spherical, namely:
IX 0

0 IX
0 0

0
I X

where I X is the common eigenvalue.


The existence of principal axes of I follows from this proposition.

(3-38)

286

5.3

Chapter 3
INERTIA ELLIPSOID AND PRINCIPAL AXES

The reduction of the equation of a quadric to its canonical form is a well-known


problem in analytical geometry (diagonalization).
We recall that the inertia ellipsoid (at o) is the locus of the extremity of the vector along every
axis through o and of norm 1 I .

At a given point o, the inertia ellipsoid shows (at least) three orthogonal axes of symmetry:
the principal axes of inertia of S (at point o).
The canonical equation of the inertia ellipsoid with respect to the system of its principal axes
oXYZ is
I X X 2 + IY Y 2 + I Z Z 2 = 1 .

(3-39)

We specify that the distance from o to some point p of the ellipsoid, namely op = 1
1 I X , 1 IY , 1
of the ellipsoid.

I is

I Z for the respective principal X,Y,Z-axes. These values are the semi-axes

(i) We consider the general case where I X I Y I Z .

If for instance I X > I Y > I Z ( 1


axis), we have:

I X is thus the semi-minor axis and 1

IZ

the semi-major

I X (cos 2 + cos 2 + cos 2 ) I X cos 2 + I Y cos 2 + I Z cos 2

that is

I X I .
The equality takes place if
I X cos 2 = I Y cos 2 ,

I X cos 2 = I Z cos 2 ,

that is if cos = cos = 0 , that is the X-axis.


Therefore, I shows a strict maximum I X for the axis corresponding to the semi-minor axis.
In the same manner,
I Z (cos 2 + cos 2 + cos 2 ) I X cos 2 + I Y cos 2 + I Z cos 2

that is

IZ I
and I shows a strict minimum I Z for the axis corresponding to the semi-major axis.

(ii) If two principal moments of inertia are equal, for example I X = I Y , then the equation of
the quadric (at o):

Mass Geometry, Inertia Tensor

287

I X (X 2 + Y 2) + IZ Z 2 = 1

is the one of an ellipsoid of revolution about the Z-axis if I Z 0 , and the one of a cylinder if
IZ = 0.

Every axis through o orthogonal to the Z-axis is a principal axis of inertia.


Finally, if the three principal moments of inertia are equal I X = I Y = I Z , then the inertia
ellipsoid (at o) is the sphere of equation
X 2 +Y 2 + Z2 =

1
IX

Every axis through o is a principal axis of inertia.


To conclude, we mention that the inertia ellipsoid (at o) is determined from the knowledge of
the principal axes and the calculation of the principal axes of inertia. Next, every moment of
inertia about any axis through o is known from op = 1 I .
On the other hand, if the principal axes are unknown, it is necessary to calculate the six
moments and products of inertia which are the coefficients of the equation of the quadric with
respect to a coordinate system. Next, the equation of the quadric is reduced to canonical form
(with respect to the system of principal axes).

5.4

MATERIAL SYMMETRIES

We recall that a material symmetry means a geometric symmetry as well as a mass


distribution one.
This well-known notion signifies, for instance, that a density is the same at two symmetric
points or that two symmetric points have the same mass.
In this case, the center of mass coincides with the center of symmetry. Indeed, given two
particles ( x1 , m) and ( x 2 , m) , we have immediately:
x1G =

1
2m

(m x1 x1 + m x1 x 2 ) =

1
2

x1 x 2 .

In the same manner, if a system has an axis of material symmetry or an axial symmetry (that
is, the material system is symmetric under rotations about an axis), then the center of mass
belongs to the axis.
Likewise, if a system has a plane of material symmetry, then the center of mass belongs to it.
For instance, for a material domain D of a (x,y)-plane of material symmetry, we have:

D z d = 0
since ( x, y, z ) = ( x, y, z ).

288

Chapter 3

We emphasize the following.


PR11 Every plane of symmetry of a material system is orthogonal to a principal axis.
Every axis of symmetry of a material system is a principal axis and the plane of other
principal axes is orthogonal to this axis.
Proof. First, we choose the (x,y)-plane as the plane of symmetry and we have thus
( x, y, z ) = ( x, y, z ).
We deduce immediately that:
Pxz = Pyz = 0

and the corresponding inertia tensor is


I x Pxy 0

I = Pxy I y 0 .
0 0 I
z

But a rotation about the z-axis makes this tensor diagonal and thus the principal axis is really
orthogonal to the plane of symmetry corresponding to the other principal axes.
In another manner, the characteristic equation is written
(I z )

Ix

Pxy

Pxy

Iy

= 0,

which shows that I z is a root of this equation and thus an eigenvalue of the inertia tensor.
Therefore, the z-axis associated with this value I z is really a principal axis of inertia.
Second, in an analogous manner, the z-axis is principal and the two other principal axes
belong to the (x,y)-plane perpendicular to the axis of symmetry.

6. STEINERS THEOREM
First, we say:
D

Principal axes of inertia of a material system S with respect to the center of mass G
are called central axes of inertia.
The inertia ellipsoid of S at G is called the central ellipsoid of inertia.

We are going to prove that the knowledge of central moments of inertia (that is at G)
determines the inertia tensor at every point.
It is obvious that the moments and products of inertia vary from one to another point of the
system and Steiners theorem is relevant to the variation of moments of inertia when parallel
axes are considered.

289

Mass Geometry, Inertia Tensor

Given an orthonormal basis, we denote by { o; xi } the corresponding frame of reference.


We consider another system of axes { G; X i } parallel to the previous and we set out to
calculate the second-order moments at point o in function of the ones at G.

Fig. 71
Let xi(G ) denote the (fixed) coordinates of G with respect to { o; xi }.
The coordinates of any point with respect to the two frames are such that

xi = xi(G ) + X i .
Before Steiners theorem, we prove the following general formula:


I (pqo ) = I (pqG ) + M ( pq d 2 x (pG ) x q(G ) )

(3-40)

where I (pqo ) and I (pqG ) denote the components of the inertia tensor of S with respect to the
respective coordinate systems { o; xi } and { G; X i }, M the mass of S and d = oG .
Indeed, we have successively:
3

I (pqo ) = [ pq ( xi ) 2 x p xq ] dm
S

i =1

= [ pq ( xi(G ) + X i ) 2 ( x (pG ) + X p )( x q(G ) + X q )] dm


S

= [ pq ( X i ) 2 X p X q ] dm + M pq ( xi(G ) ) 2 Mx (pG ) xq(G )


S

+ 2 pq
i

xi(G )

S X i dm

x (pG )

S X q dm xq S X p dm .
(G )

The three last integrals vanish since they are just the components of M X (G ) with respect to
{ G; X i } which obviously vanish.

290

Chapter 3

Thus we have :
I (pqo ) = [ pq ( X i ) 2 X p X q ] dm + M pq d 2 M x (pG ) x q(G ) .
S

The previous integral represents every moment of second order relative to the frame { G; X i }
of which the axes are parallel to the corresponding ones of { o; xi } and we have obtained
consequently:
I (pqo ) = I (pqG ) + M ( pq d 2 x (pG ) x q(G ) ) .
Now we can state the Steiners theorem.
PR12 The moment of inertia about any given axis is the moment of inertia about the
parallel central axis C plus the moment of inertia about the given axis if all the mass
were located at the center of mass.
It is expressed as


I = IC + M d 2

(3-41)

where d is the distance between o and G.


Proof. The moments and products of inertia about any axis parallel to the corresponding
central axis are easily obtained from the ones about the central axis.
For instance, if p = q = 1 :
(o)
(G )
(G )
I11
= I11
+ M (d 2 x1(G ) x1(G ) ) = I11
+ M ( yG2 + xG2 )

where yG2 + z G2 is the square of the distance between the parallel axes ox and GX.
Since there is such an equality for every coordinate axis, we have proved (3-41).
The following is obvious:
PR13 Among the moments of inertia relative to a set of parallel axes, the smallest is the one
about the central axis.
Example. The moment of a thin homogeneous rod S of length 2l about a central axis C
perpendicular to the rod is
l

l2

I C = x 2 dm = x 2 dx = M
S

Steiners theorem confirms this result since, given the axis through the origin of the rod
and parallel to C, we have:
2l

I = x 2 dx =

Ml2

and thus
IC = I M l 2 = M

l2
3

Mass Geometry, Inertia Tensor

291

7. EXERCISES
We consider only a few exercises because the inertia tensor notion is developed in
intermediate courses in mechanics at the undergraduate level (in the matrix context).
Exercise 1.
Given the symmetric tensor

7 3
3 1

0 0

0
0
8

with respect to an orthonormal frame {o;x,y,z}, find the principal axes and the orthogonal
transformation from the initial frame to the one of principal axes.
Answer. The characteristic equation is immediately
( 8)(2 6 16) = 0

and the eigenvalues are the double root 1 = 2 = 8 and 3 = 2 .

For the double eigenvalue, every eigenvector v = vi e i is obtained by solving the system
v1 3v2 = 0 ,

3v1 9v 2 = 0 ,
0 = 0

and from this system, we deduce v1 = 3v 2 while v3 is arbitrary.


For this double eigenvalue there is obviously a two-parameter family of possible
eigenvectors:
v = 3v 2 e1 + v 2 e 2 + v3 e3 .

Within this family of eigenvectors, first we choose the unit eigenvector e 3 denoted E 3 and
second we choose the unit eigenvector

3
10

e1 +

1
10

e 2 denoted E 2 .

We point out that these orthogonal vectors are really eigenvectors associated with = 8 since
we have directly:
I E2 = 8 E2 .

I E3 = 8 E3

For the eigenvalue 3 = 2 , every eigenvector v = vi e i is obtained by solving the system


9v1 3v 2 = 0 ,

3v1 + v2 = 0 ,
10v = 0
3

v 2 = 3v1 ,

v 3 = 0 .

292

Chapter 3

Within the one-parameter family of possible eigenvectors, we choose the unit eigenvector
E1 =

1
10

e1 +

3
10

e2

so as to obtain a direct basis ( E1 , E 2 , E 3 ) or right-handed coordinate system oXYZ.


We point out this vector E1 is really an eigenvector corresponding to the eigenvalue 3 = 2
since we have immediately

I E 1 = 2 E 1 .
In conclusion, the three unit vectors E1 , E 2 , E 3 define principal axes of I.
In addition, let us prove that PR9 is verified; namely:
Every unit vector perpendicular to the principal vector E1 ( 3 = 2 ) is an eigenvector
corresponding to 1 = 2 = 8.
Indeed, E 3 = e 3 as well as E 2 =

3
10

e1 +

1
10

e 2 are perpendicular to E1 and thus any vector

of the plane determined by E 2 and E 3 is so. Such a unit vector is written a, b R :


3a
10a 2 + b 2

e1 +

a
10a 2 + b 2

e 2 + b e3 .

It is really an eigenvector corresponding to the double eigenvalue 8 since


2
2
2
2

7 3 0 3a 10a + b 3a 10a + b

2
2
2
2
3 1 0 a 10a + b = 8 a 10a + b
.
0 0 8

b
b

Finally, the orthonormal basis (e1 , e 2 , e 3 ) of the initial frame oxyz is transformed into the
orthonormal basis ( E1 , E 2 , E 3 ) of the frame oXYZ of principal axes. The corresponding
orthogonal transformation is defined by
E1 1 10 3 10 0 e1


E 2 = 3 10 1 10 0 e 2 .

E 0
0
1 e 3
3

Exercise 2.

Given a material system S determined by the illustrated homogeneous tetrahedron (of


density ), find the inertia tensor of S at o,
(i) with respect to the frame oxyz,

293

Mass Geometry, Inertia Tensor

(ii)

with respect to the frame oXy z , the y -axis being along the bisector of the angle
(xoy ) and X axis perpendicular to the previous.

Answer.

Fig. 72
(i) Since the element of mass of the tetrahedron is
x2
dm = d =
dz
2

and since
x z
+
=1
a 2a

x = a z 2,

we deduce that the mass of the tetrahedron is


M =

2a

2 0

z
(a ) 2 dz = a 3 .
2

We note that the axes of the frame oxyz are not principal axes and thus we will have to
calculate the products of inertia as follows.
We note that there is a simplification since
Ix = Iy,

Pxz = Pyz .

First, let us calculate:

I x = ( y 2 + z 2 ) dm = I xoy + I zox .
S

From
I zox = y 2 dm = x 2 dm =
S

and
I xoy = z 2 dm =
S

we deduce:

2a

2 0

2a

2 0

3
z
(a ) 4 dz = a 5 = Ma 2
2

2
z
z 2 (a ) 2 dz = Ma 2
2

294

Chapter 3

I x = Ma 2
and also
I z = ( x 2 + y 2 ) dm = 2 x 2 dm =
S

6
5

Ma 2 .

Now, we calculate the products of inertia:


2a

a z 2

Pxy = xy dm = dz
S

a5

40

40

19

a x

xy dm =

2a

2 0

dz

2a

z dz

a z 2

x(a x) 2 dx

Ma 2
2a

a z 2

Pyz = yz dm = dz
=

dx

dx

a x

yz dy =

2 0

a z 2

(a x) 2 dx

Ma 2

80

The inertia tensor with respect to oxyz is thus written:

3 40

19 80

I = Ma 3 40
1
19 80 .

19 80 19 80 6 5
It is better to take the symmetry into account and to use the frame with the y -axis.

(ii) The X-axis is a principal axis of inertia since it is orthogonal to the ( y ' , z )-plane which is
a plane of symmetry.
Let us determine the inertia tensor at o with respect to oXy z .
The moment of inertia about the X-axis is obtained from (3-31), that is (since cos = 2 2 ,
cos = 2 2 , cos = 0 ):
IX =

1
2

43

40

I x + I y + Pxy =

Ma 2 .

The moment of inertia about the y -axis is (since cos = 2 2 , cos = 2 2 , cos = 0 ):
I y' =

1
2

37

40

I x + I y Pxy =

Ma 2 .

Now, to find the product of inertia Py ' z , we calculate the moment of inertia about the b-axis
along the bisector of the angle ( y ' oz ) in two ways.
First, with respect to oxyz, since cos = 1 2 , cos = 1 2 , cos = 2 2 , we have:

295

Mass Geometry, Inertia Tensor

Ib =

1
4

=(

19 2

) Ma 2 .

I x + I y + I z Pxy

85
80

80

2
2

Pyz

2
2

Pzx

Second, with respect to oXy' z , since cos = 0 , cos = 2 2 , cos = 2 2 , we have:


Ib =
=

1
2

I y ' + I z Py ' z
2

85
80

Ma 2 Py ' z .

The comparison between the two results implies:


Py ' z =

19 2
80

Ma 2 .

Since the ( y ' , z ) -plane, orthogonal to the principal X-axis, is a plane of symmetry, we have:
PX y ' = PX z = 0

and the tensor of inertia with respect to oXy ' z is written:


43
40

37
19 2
Ma 2 .

40
80

19 2
6

80
5
0

From the corresponding characteristic equation, the reader will calculate the two last principal
axes of inertia.

Exercise 3.

As shown in Fig. 73, a material system S is made up of particles of


-

a conic system S1 of height h, radius R and constant density 1 ,


a cylindrical system S 2 of height H, radius R and constant density 2 .

(i) Calculate the distance between the origin o and the center of mass G of S,
(ii) Find the central tensor.

296

Chapter 3

Answer.

Fig. 73
(i) The height of the center of mass of the cone is easily obtained as follows.
The mass of the conic system S1 being M 1 , we have:

M 1 z G1 = z dm = 1
S1

= 1 2 r
0

h2
2

d r dr

(1

r
R

z dz

2
r2
2 R

)
dr
=
h
1
4
R2

= M1 h.
4

So, with respect to (e1 , e 2 , e 3 ) , the center of mass G1 is such that


oG1 =

3
4

h e3

and the center of mass G2 of S 2 is such that


oG 2 = (h +

H
) e3 .
2

From

( M 1 + M 2 ) oG = M 1 oG1 + M 2 oG 2
we deduce that
oG =

1
H
3
( M 1 h + M 2 (h + )) .
M1 + M 2
4
2

(ii) The axisymmetric system about the z-axis is a principal axis and we know that I x = I y .

First, we calculate the inertia moment of S1 about the z-axis:

297

Mass Geometry, Inertia Tensor

I z(1)

= ( x + y ) dm = 1
2

S1

3
10

h
hr
R

d r dr

dz = 1

R4h
10

M1R 2 .

Likewise, the inertia moment of S 2 about the z-axis is


I z( 2) = ( x 2 + y 2 )dm = 2
S2

R 3

0 r

dr dz =

2 R 4 H

= M 2R2.
2

In conclusion, the moment of inertia of S about the x-axis is


R2

I z = I z(1) + I z( 2) = (3M 1 + 5M 2 )

10

Next, we calculate the moment of inertia of S1 about the x-axis:


I x(1) = ( y 2 + z 2 ) dm =
S1

(since I x(1) = I y(1)

1 (1)
Iz
2

+ z 2 dm
S1

dm = y 2 dm ).

From

z 2 dm = 1

d r dr h r z 2 dz = 1 R 2 h 3
0

= M 1h 2 ,
5

we deduce:
I x(1) = I y(1) =

3
20

M 1 ( R 2 + 4h 2 ) .

In the same manner, the moment of inertia of S 2 about the central X-axis (parallel to x-axis)
is such that
I X( 2) =

1 ( 2)
Iz +
z 2 dm .
S
2
2

Since
2

I z( 2) = ( x 2 + y 2 ) dm = d
S2

r 3 dr

H 2

H 2

dz = M 2

and

S
we deduce:

z dm = 2
2

r dr

H 2

H 2 z

dz = M 2

H2
12

R2
2

298

Chapter 3

I X( 2)

= M2(

R2
4

H2
12

).

Steiners theorem leads to the moment of inertia of S 2 about the x-axis:


I x( 2) = M 2 [

R2
4

H2
12

+ (h +

H
2

)2 ] .

Therefore, the moment of inertia of the system S about the x-axis is


I x = I x(1) + I x( 2) =

3
20

M 1 ( R 2 + 4h 2 ) + M 2 [

In conclusion, the central tensor (that is at G) is

IX 0 0

0 IX 0
0 0 I
Z

where

I X = I x ( M 1 + M 2 ) oG
and I Z = I z obviously.

R2
4

H2
12

+ (h +

H
2

)2 ] = I y .

CHAPTER 4

KINETICS AND DYNAMICS OF SYSTEMS

Kinetics, which is elaborated from kinematics and mass geometry, is dealt with
notions of momentum, angular momentum and kinetic energy.
Dynamics analyzes the relationships between motions and mechanical actions that are
the forces.
From works of Stevinus, Galileo, Kepler and Huygens, Newton stated the fundamental
laws of classical mechanics in his Philosophiae Naturalis Principia Mathematica (1687).
Later, dAlembert, Euler, Lagrange, Laplace, Poinsot, Poisson, Jacobi, Hamilton,
Poincar (and others) made mechanics progress in connection with astronomy notably.
However, atomic or galactic phenomena with velocities approaching the speed of
light, enormous masses, etc. cannot be described in classical mechanics and thus require new
mechanics.
But innumerable applications of engineering and astronautics for instance reveal the
importance of classical dynamics nowadays. The following postulates form the axiomatics of
this branch.

1. NEWTONS POSTULATES
We recall that the frame of classical mechanics is an affine Euclidean space, which is
homogeneous (no privileged point a priori) and isotropic (no privileged direction a priori).
The position of every particle of a material system S is a point of this space at a given time.
A corresponding vector space is defined by relating any position to a reference point
(observer at the same time); that is, by choosing an origin of the space.
Let E be a Euclidean point space,
E be the vector space associated with E,
R be a frame of reference,
be an open of E,
I be an interval of R.

299

300

Chapter 4

Given a system S of N particles p h defined by N points x1 ,..., x N of respective position


vectors ox h (t ) = rh (t ) and respective masses m1 ,..., m N , we say:
D

A motion of S is a set of N vector functions of class C 2


rh : J ( I ) : t  rh (t ) = x h
such that:
t J , h k : x h x k .

A force acting on p h (with respect to R ) is a function of class C 1

f h : I N E N E : (t , x1 ,..., x N , x 1 ,..., x N )  f h (t , x1 ,..., x N ) .


A material domain of N particles is so defined.
Notation. The velocity of any xh with respect to R is denoted by v h .

1.1

EXPERIMENTAL LAWS

Let r( q ) (t ) denote the position vector of a particle (or body), at the time t, of a qth
experiment.
We consider the three following experimental phenomena.
(i)
A body is fastened to the end of a vertical spring of which the other extremity is fixed.
If the body has various vertical motions, we observe that the acceleration is always
proportional to the elongation.
With another body, we observe the same experimental law but the acceleration is multiplied
by another coefficient.
For every experiment, this law is expressed as follows:
m r( q ) (t ) = k r( q ) (t )

where the positive coefficient m, peculiar to each body, is called the mass of the body and the
constant k concerning the spring is called the stiffness.
An inductive reasoning, which postulates that such experimental results will be always
obtained, leads to the following conclusion:
A motion
r : I ( R) E : t  r (t ) ,

with initial conditions r (t 0 ) = r0 and r(t 0 ) = v 0 (given r0 and v 0 ), mathematically verifies


the differential equation:
m r = k r .

Kinetics and Dynamics of Systems

301

(ii)
At the beginning of the 17th, Galileo which made observations concerning bodies in
free fall and motions on inclined planes concluded that heavy bodies fall with the same
acceleration.
For every experiment, this law is expressed with the gravity acceleration g as follows:
r( q ) (t ) = g

An inductive reasoning leads to a theoretical law given by the differential equation:


m r = m g .

(iii) From Keplers laws (deduced from Tycho Brahes observations of planetary motions),
Newton found an experimental law which holds for all the planets of the solar system:
r( q ) (t )

r( q ) (t ) = k

r( q ) (t )

where k is a constant of proportionality. This is valid in the frame of reference with origin at
the center of mass of the solar system and axes in the directions of three fixed stars.
An inductive reasoning leads to the following differential equation
m r = k m

1r
r

To sum up, the three previous experiments lead to a common experimental law, namely:
Given the observed motions, there is a vector function f such that
m r( q ) (t ) = f (m, r( q ) (t ))

where the vectors are referred to a well-determined frame of reference (to be specified).
An inductive reasoning introduces in all the cases a differential equation of type:
mr(t ) = f (m, r (t ))

with the initial conditions r (t 0 ) = r0 and r(t 0 ) = v 0 .


In other experiments, the velocity-vector r(t ) can occur and time can explicitly appear too.
Therefore, the differential equation ensuing from experiments is written (with respect to R ):
m r(t ) = f (t , r (t ), r(t ))

(4-1)

where the coefficient m is removed from f since it is constant during the motion.

1.2

POSTULATES

(i)

Postulate of Initial Conditions

The conclusions of works of Galileo (1638) can be stated as follows:


PO

The position vector of any particle is determined at every time if the position and the
velocity of the particle are known at initial time.

302

Chapter 4

So, the acceleration vector of the particle is determined at every time too.

(ii)

Newtons Second Law (Motion Law)

In a way, Newton introduced the notion of momentum.


Newtons second law also called the fundamental principle of dynamics can be stated as
follows:
PO

The time derivative of the linear momentum of every particle is collinear to the
resultant force acting on the particle in an absolute frame of reference.1

We denote:
f =

d
(m r ) .
dt

(4-2)

If the mass is constant, this postulate expresses that the differential equation
f (t, r(t), r(t)) = m r(t)

must be verified for every motion defined by


r : I : t  r (t ) .

The coefficient of proportionality between the force f and the acceleration r defines the mass.
This law is valid in relativistic mechanics.
(iii)

Newtons First Law (Inertia Law)

The previous motion law of Newton has the following corollary also called the law of inertia,
namely:
PO

In the absence of forces acting on a particle, its linear momentum is constant.

We denote:
f =0

m r = c

(constant vector c).

(4-3)

This law is valid in relativistic mechanics.


If the mass is invariable, this law is written:
f =0

r = k

(constant vector k)

and it is the famous Galilean principle of inertia:


PO

A particle remains at rest or continues to move in a straight line with a uniform


velocity if no force acts on it.

With this principle, Galileo puts an end to the false Aristotelian principle which claimed that
every motion (even uniform rectilinear motions) must be caused.
1

This well-known notion is recalled later.

Kinetics and Dynamics of Systems

303

(iv)

Newtons Third Law (Principle of Action and Reaction)

PO

The forces of action and reaction between interacting particles are collinear, opposite
in direction and equal in norm.

Explicitly, if f hk denotes the force exerted by the particle p k = ( x k , mk ) on the particle


p h = ( x h , mh ) , we have:
x k x h , c hk R+ :

f hk = c hk x h x k ,
f hk = f kh .

(4-4)

This principle is not valid in relativistic mechanics.


(v)

Law of Parallelogram of Forces

Stevinus introduced the well-known law of parallelogram of forces in statics (1586) and
Newton expressed it in dynamics later.
It is a corollary for Newton:
CO

If several forces act on a particle they are combined according to the parallelogram
law of addition of forces.

1.3

GALILEAN RELATIVITY AND INERTIAL FRAMES

The postulates of dynamics are implicitly referred to a frame given the presence of
velocities and accelerations. More precisely, the fundamental law of dynamics of N particles:
mh rh = f h (t , r1 ,..., rN , r1 ,..., rN )

h = 1,..., N

defines a motion of the material system corresponding to the 2N initial conditions:

rh (t 0 ) = rh0 ,

rh (t 0 ) = v h0 ,

given rh0 and v h0 at t 0 .


If this law holds in a given frame, then it will not be verified in another frame accelerated
with respect to the previous one.
An essential problem is to find a class of frames of reference for which the three Newtons
laws hold.
D

Every frame of reference such that Newtons laws hold is called an absolute frame of
reference.

This cannot be checked directly, this can be made by calculating the motions from the
postulates and next by comparing with experimental results.

304

Chapter 4

So a local reference frame attached to the surface of the earth; that is, a trihedron with axes in
the apparent vertical, south and east directions, is suitable for most problems of classical
mechanics. Time is measured from the rotation of earth (sideral time).
But there are discrepancies between calculations from postulates and observations. For
example, there exists deviations of landing places of falling bodies, of rockets and so on.
Discrepancies between observations and theoretical predictions vanish if a heliocentric
trihedron is chosen; that is, such that its origin is the center of mass of the solar system and
the directions of axes are fixed with respect to the stars. This heliocentric trihedron, also
called the trihedron of Copernicus, and the sideral time explain almost all the phenomena.
However, for instance, we observe that the moon is early with respect to the calculated time.
This anomaly follows from the variation of the rotation of earth (tides, etc.) and vanishes if
the atomic clock is chosen. The atomic time is more accurate than astronomical time
(earths period); it corresponds to the period of an atomic vibration (see physics and
astronomy).
We conclude:
PR1

The truth of the postulates of classical mechanics was always held, but successive
changes of frames of reference were necessary.
The trihedron of Copernicus and atomic time define an absolute frame of reference.

Now let us recall the essential notion of inertial frame.


Let

R = { o; e , e , e } denote an absolute frame of reference.


1

Question. How is the expression of the fundamental principle of classical dynamics in an


orthonormal frame R E = { O(t ); E1 (t ), E 2 (t ), E 3 (t ) } moving with respect to R ?
We assume that the functions O(t ) and E i (t ) are of class C 2 .

We know that the acceleration of any point x in R (called the absolute acceleration) is
written:
a A = a R + aT + aC
where the relative acceleration a R is the acceleration of x with respect to the moving frame
R E , a T denotes the transport acceleration of x and a C is the well-known Coriolis
acceleration a C = 2 v R where v R is the relative velocity of x.
If f A denotes the resultant of forces acting on x for an observer of R, then the fundamental
principle of dynamics, namely:
f

is written:


= ma A

maR = f

m aT m aC .

(4-5)

D  A frame of reference is said to be inertial if the fundamental principle of dynamics has


the same form as in the frame of Copernicus (with the same clock):


R E inertial

iff

maR = f A .

Kinetics and Dynamics of Systems

305

There is a necessary and sufficient condition for inertial frame existence which is called the
principle of Galilean relativity.
More precisely, let us prove the following:
PR2

A frame is inertial iff it moves with constant velocity vector with respect to the
Copernicus frame R (uniform rectilinear translation).

Proof. First, the hypothesis


maR = f

implies

aT + aC = 0

d 2 oO
dt 2

(since m 0 )

d
dt

Ox + ( Ox ) + 2 v R = 0 .

This equation being verified for any point x, in particular for O, we have:
Ox = v = 0

d 2 oO
dt 2

= 0.

Thus, we have for any x:


d
dt

Ox + ( Ox ) + 2 v R = 0 .

This equation being verified for any point x, in particular for x such that Ox = E1 , we have
d
v R = 0 in this case; thus, by letting  =
R , this equation becomes:
dt

 E1 + ( E1 ) = 0
that is:
E1

E2

E3

 3 + (. E1 ) E1 = 0 .

By considering the components of the vectors of this equations with respect to the moving
basis of R E , the resulting system:
0 + 1 1

= 0,

 3 + 1 2 = 0 ,
 2 + 1 3 = 0

has the following solution:

( 1 ) 2 =

2 =3 = 0.

306

Chapter 4

In the same manner, if x is such that Ox = E 2 in particular, we obtain 1 = 0 .


Thus the various results lead to:

= 0.
In conclusion, a moving inertial frame R E = { O; E1 , E 2 , E 3 } is such that:
d 2 oO
dt 2

=0

=0,

which means that R E moves in translation with constant velocity vector with respect to R,
it is said to be in uniform rectilinear translation.
Conversely, if the motion of R E is a uniform rectilinear translation with respect to a frame
of Copernicus, that is:
d 2 oO
dt 2

=0

= 0,

then we deduce:
aT = a C = 0 .

Therefore, the frame R E is inertial since the fundamental principle of dynamics is written as
in the frame of Copernicus, namely:
maR = f a .

From the previous proposition, we conclude:


PR3

All the inertial frames (which are in uniform rectilinear translation) are equivalent.

Inertia law (or Newtons first law) being held, no unique frame can be privileged and a
class of inertial frames is so defined.
In other words, a frame moving in uniform rectilinear translation with respect to an absolute
frame of reference is absolute. The conclusions of classical mechanics are the same in both
frames.
So, for various observers of different absolute frames, the force exerted on any particle is
invariable and we recall:
PR4

The laws of classical mechanics are invariable through any change of inertial frame.

Remark. Concerning noninertial frames of reference, the principle of action and reaction is
not valid!

Indeed, given any pair (a,b) of isolated mass points, this principle
f ab + f ba = 0

does not hold with respect to a noninertial frame since, in such a frame, the forces are
expressed as
Fab = f ab f T f C ,
Fba = f ba f T f C

Kinetics and Dynamics of Systems

307

and thus we have:

Fab + Fba 0 .
In practice, noninertial frames of reference are often used provided the forces of transport and
of Coriolis can be neglected. This is not always possible as the pendulum and gyroscope of
Foucault prove it.

2. KINETICS
Let us introduce the well-known notions of linear momentum and angular momentum,
elements of a dynam, as well as the one of kinetic energy.

Given an inertial frame of reference


= { o; e1 , e 2 , e3 }, we denote by v h the velocity of any
particle ( x h , mh ) of a system S of N particles.

2.1

KINETIC DYNAM

The linear momentum of S with respect to R is the vector sum of momenta of all
particles:
N

P = mh v h

[or

h =1

S v ( x) dm ] .

(4-6)

By denoting any position vector rh = ox h , we say:


D

The angular momentum of S about o with respect to R is the vector sum of momenta
of all particles:
N

Lo = rh mh v h

[or

h =1

S r v ( x) dm ] .

The kinetic dynam of S with respect to R is defined by its following elements of


reduction at point o:
P
L .
o

Remark. We really have:

La = Lb + ab P
since

(4-7)

ax h mh v h = ab mh v h + bx h mh v h .
h

308

Chapter 4

Angular momentum expression

From
Lo = rh mh
h

drh
d
(oG + Gx h )
= (oG + Gx h ) mh
dt
dt
h
d
d
Gx h + Gx h mh v G + Gx h mh Gx h
dt
dt
h
h

= oG mh v G + oG mh
h

and because (invariable masses):


oG

d
( mh Gx h ) = 0 ,
dt h

mh Gx h v G = 0 ,
h

we deduce:
Lo = oG M v G + Gx h mh (v h v G ) .

(4-8)

Koenig expression of angular momentum

Let R e = { G; e1 (t ), e 2 (t ), e 3 (t ) } be a frame in translation with respect to R and the origin of


which is the moving center of mass G of S.
PR5

The angular momentum of S about o is the moment about o of the linear momentum
of the total mass concentrated at G plus the angular momentum, about G, of S in
relative motion with respect to R e :
N

Lo = oG M v G + Gx h mh v hR .

(4-9)

h =1

The expression in relative motion with respect to R e means that the angular momentum is
expressed in function of velocities v hR relative to R e .
Proof. The velocity v h of any point xh with respect to R is

v h = v G + v hR
since the transport velocity is the velocity of the origin of R e in this particular case.
From (4-8) we deduce the announced expression (4-9)

2.2

KINETIC ENERGY

The kinetic energy of a system S with respect to R is the sum of kinetic energies of
all particles:
T=

1 N

h =1

mh v h2

(or T =

v
2 S

( x) dm ).

(4-10)

309

Kinetics and Dynamics of Systems

Koenig expression of kinetic energy


We consider again the translating frame R 'e .
PR6

The kinetic energy of S equals the energy of mass-center translation of S as a whole


plus the energy due to motion of all particles relative to the center of mass:
T=

1
2

M v G2 +

mh (v hR ) 2 .

(4-11)

h =1

Proof. Kinetic energy is written:


T=

mh [ (oG + Gx h )]2

2
dt
h

mh [(
2
h

dGx h 2
doG 2
doG dGx h

+(
) +2
) ].
dt
dt
dt
dt

The second term of the sum vanishes since:

mh
h

doG dGx h doG d

=
mh Gx h = 0
dt
dt
dt dt h

and the proposition is thus proved if we recall that the velocity of any xh relative to the
translated frame R e is denoted v hR .

3. THEOREMS OF MECHANICS OF SYSTEMS


Theorems of linear momentum, angular momentum and kinetic energy follow from
the postulates of classical mechanics.
In an inertial frame of reference

3.1

R = { o; e , e , e } we consider a system S of N particles.


1

FIRST INTEGRALS OF A SYSTEM OF PARTICLES

The exact description of individual motions of N particles in gravitational interaction


is generally impossible. The three-body problem cannot be already solved in the general case.
The N bodies obey to the Newtons law of gravitation. We recall that the long-range forces of
attraction of gravity cannot be removed (for example with the help of a screen as for
electromagnetic interactions).
The N-body problem consists in determining the position and velocity of every particle, at
each instant, from initial conditions that are the position and velocity of every particle at a
given time t 0 .

310

Chapter 4

This problem is fundamental in planetary astronomy, celestial mechanics, stellar dynamics,


plasma physics for instance. Let us show its equations.
In the inertial frame of reference R, we consider N points xh of respective masses mh and
coordinates ( x h , y h , z h ) .
Each particle p h is attracted by the N 1 other particles with a force:
fh =

mh mk
rhk2

k =1
k h

(4-12)

1hk

where

1hk =

rhk
rhk

( rhk = x h x k ).

We have thus N differential equations:


mh

d 2 rh
dt

k =1
k h

mh mk
rhk3

rhk .

Projections onto the coordinate axes lead to 3N differential equations, namely:


mh

d 2 xh
dt

G mh mk ( x k x h )

k =1
k h

(( x k x h ) + (( y k y h ) 2 + ( z k z h ) 2 ) 3 2
2

and similar others in y and z.


So, we have obtained a system of 3N second-order differential equations or 6N first-order
equations to be solved with 6N unknowns which are the positions and velocities of particles.
Therefore, 6N integrations are necessary, unless we find first integrals. We recall the
following
D

A first integral of motion equations is a function of time, positions and velocities


which remains constant during the motion.

3.2

LINEAR MOMENTUM THEOREMS

3.2.1 Linear Momentum Theorem


PR7

Given an inertial frame, the time rate of change of linear momentum of S equals the
resultant of external forces on S:
dP
= f (e ) .
dt

Proof. With respect to an inertial frame, we have:

(4-13)

Kinetics and Dynamics of Systems

311

dP
d
= (mh v h )
dt
h dt

and from the fundamental principle of dynamics (Newtons second law) we deduce:
dP
= f h(e )
dt
h

where only the external forces must be taken into account since we assume that the internal
forces act along the line joining every pair of particles and are equal in norm and opposite in
direction (Newtons third law).
The theorem is thus proved.
Remark. We note that the impulse delivered by the resultant of external forces is the
following increase:
P (t ) P (t 0 ) =

t0

f ( e ) dt .

(4-14)

3.2.2 Theorem of Conservation of Mass


PR8

In the absence of external forces, the linear momentum of S remains unchanged in an


inertial frame.

Proof. The implication


f (e ) = 0

d
dt

mh v h = 0
h

leads to the following first-order differential equation:


P = mh v h = c

(constant vector).

(4-15)

Example. A gun recoils as soon as the explosion occurs since only internal forces play a role
and the system weapon-projectile is initially at rest ( c = 0 ).

3.2.3 Theorem of Motion of Mass Center


First, the linear momentum of a system S of constant masses is the product of the
mass of S and the velocity of mass center:1


P = M vG

(4-16)

since
P = mh v h =
h

We denote

v (G ) by v G .

d
d
( mh rh ) = ( M oG ) = M v G .
dt h
dt

312

Chapter 4

Second, we say:
PR9

In an inertial frame, the center of mass of S moves as a particle whose mass is the one
of S and on which acts a force equal to the resultant of external forces exerted on S:

dv G
= f (e ) .
dt

Proof. This is obvious from the equation M

(4-17)
dv G dP
and PR7.
=
dt
dt

Remark. Since the internal forces do not act, this theorem is interesting to study complicated
systems as the human body for instance.
Examples. The center of mass of an exploded space capsule continues to follow its orbit
since the explosion gives rise to only internal forces.
In the same manner, internal forces lead the driving wheels of a vehicle (at rest) to spin on the
ice, since the center of mass must remain at rest. The moving off is possible if there are
external friction forces (opposed to the rotation forces).

3.2.4

Special Case of Rigid Bodies

We consider a rigid body B which is free to translate or rotate with respect to an


inertial frame of reference R.
Let R E = { O(t ); E1 (t ), E 2 (t ), E 3 (t ) } be a frame fixed in the body where O B and the
various E i depend on time (with respect to R ).
Notation. Unless otherwise specified, the times derivatives (and thus velocities) are related to
the inertial frame of reference R.
Angular velocity tensor
We know that the various vectors

dE i
can be expressed with respect to the moving basis
dt

( E i ) as
dE i
= ij E j
dt

i, j = 1,...,3.

PR10 The components ij are the components of a


tensor, denoted by .
Proof.

(4-18)

( )-tensor called the angular velocity


1
1

Given any change of bases of the moving frame (said to be a change of basis
connected to B):
E j = ij E i ,

E k = ks E s

Kinetics and Dynamics of Systems

where the various ij are constants, then the vectors


respect to the basis ( E j ) as follows:
dE j
dt

dE j
dt

313

are obviously expressed with

= jk E k .

From

jk E k =

dE j
dt

= ij

dE i
= ij ir E r = ij ir rk E k
dt

we deduce the expected result:

jk = ij rk ir .
This proposition being proved, let us make explicit the components of angular velocity tensor.

Fig. 74
Let ox be the position vector of any x B :
ox = oO + Ox

where O is a fixed point in B.


We have:

ox = oO + X i E i
where the components X i of X = Ox are fixed with respect to ( E i ) .
At time t, the velocity of point x (with respect to R) is
v ( x) = v (O ) + X i

dE i
dt

also denoted:
v x = vO + X i

dE i
,
dt

i = 1,2,3

314

Chapter 4

and thus, the components of this velocity vector relative to the basis ( E i ) of R E fixed in B
are expressed as

v j ( x) = v j (O) + ij X i
where the various ij are the components of the

(4-19)

( )-tensor of angular velocity.


1
1

The covariant components of are

ij = g hj ih = ih E h . E j
which implies


ij =

dE i
Ej
dt

(4-20)

Remark. It is obvious that no confusion is possible between the angular velocity tensor
and the symbol of components i j = ikj du k encountered in Section 5.3 of Chapter 2; in this

last case the corresponding elements ij are not constant.


PR11 Angular velocity tensor is antisymmetric.
Proof. Since E i . E j is constant (for example E i . E j = ij for an orthonormal basis), we
deduce:
dE j
dE i
d
( Ei . E j ) =
E j + Ei
=0
dt
dt
dt
that is:

ij = ji .

(4-21)

Angular velocity vector

Let us consider a basis ( E1 , E 2 , E 3 ) of a Euclidean space connected to B.

()

We know that the adjoint of the 02 -tensor of components ij is a vector, denoted , such
that
1
i=
( jk )i ( jk )
det g
and called the angular velocity vector.
We note:

jk = det g ijk i .
In Section 5.5.2 of Chapter 2 we defined the adjoint of exterior product of two vectors. By
considering formula (2-114), given vectors of components i and X j , the following
expression:

Kinetics and Dynamics of Systems


det g ijk i X

315

represents the covariant components of the vector product


Ox

(also denoted X ).

Notation. According to usage, we denote the angular velocity vector by , no confusion


with the angular velocity tensor being possible.
In conclusion, we have shown the general coordinate presentation as well as the vector
presentation of velocity field of points of a rigid body, which is really:


v ( x) = v (O) + X

(4-22)

that is:
v ( x) = v (O) + X i ( E i ) = v (O) + X i

dE i
dt

that is also written:




v ( x) = v (O) + ij X i E j

(4-23)

[cf. (4-19)].
We note that vector is an axial vector; that is, a vector whose directional sense depends on
the handedness of the frame, just as the vector product! But, the use of the angular velocity
tensor does not require any convention as regards to the frame orientation and lets use any
coordinate system..
Example. Make explicit the components of the linear momentum P in function of v G for any
rigid body.
The linear momentum is written:
P = v ( x) dm
S

= M v (O) + X dm .
S

Since
OG =

we have:

M S

X dm ,

P = M v (O) + M OG .

From the following derivative (with respect to R):


doG
= v G = v (O) + OG ,
dt

we find again:
P = M vG .

316

Chapter 4

To conclude, we emphasize that in any coordinate system, the covariant components of


OG are:
det g ijk i X Gj .
But, in an orthonormal basis of Euclidean space, contravariance and covariance are
indistinguishable (variance is indifferent) and det g = 1 ; the components of P are

P k = M (v k + ijk i X Gj ) .

(4-24a)

For instance:

P1 = M [vo1 + ( 231 2 X G3 + 321 3 X G2 )]


= M [v O1 + ( OG )1 ] .
In an equivalent manner, we have:


i
P i = M (vO
ijk X Gj k )

(4-24b)

since
i
i
P i = M (vO
+ jki j X Gk ) = M (vO
+ kji X Gj k ) .

3.3

ANGULAR MOMENTUM THEOREMS

Let us consider any system S of N mass points ( x h , mh ) .


3.3.1 Angular Momentum Theorem

PR12 In an inertial frame, the time derivative of angular momentum of S about any point a
is equal to the moment about a of the external forces plus the vector product of the
linear momentum of S and velocity of a:
N
d
La = (ax h f h( e ) ) + P v a .
dt
h =1

Proof. Given an inertial frame of reference


principle of dynamics is written:

(4-25)

R = { o; e , e , e }, we know the fundamental


1

d
(mh v h ) = f h(e ) + f h(i ) .
dt

Given any a in R, we have:


d
d
La = (ax h mh v h )
dt
h dt

= [(
h

d ox h d oa
d

) mh v h ] + [ax h (mh v h )]
dt
dt
dt
h

Kinetics and Dynamics of Systems

317

= v a mh v h + [ax h ( f h( e) + f h(i ) )]
h

= v a P + (ax h f h(e ) ) ,
h

since we assume that the internal forces act along the line joining every pair of particles and
are equal in norm and opposite in direction (Newtons third law).

3.3.2 Relation between Kinetic Dynam and Dynam of Forces

The applications for which the term P v a vanishes are essential. It is the case where
the reference point a is fixed (and thus chosen as origin o of R); then, expression (4-25)
becomes the following
d
Lo = rh f h(e )
dt
h

also written:
d
Lo = M o(e ) .
dt

(4-26a)

It is also the case where the mass center G is chosen as reference point a. Indeed, if the
masses are invariable, theorem of mass center ( P = M v G ) leads to the similar result:
d
LG = M G(e ) .
dt

(4-26b)

Therefore, we say:
PR13 In an inertial frame, the time rate of change of angular momentum of S about a fixed
point or about the mass center equals the moment about the concerned point of the
resultant of external forces acting on S.
We note that the term P v a also vanishes if v a is collinear to v G , but this case is less
interesting.
In conclusion, there is the following relationship between kinetic dynam and dynam of forces:


d P

dt L o

f e)
=
.
(e)
M o
G

Remark. Later, we will consider the notion of dynamic dynam defined as following.

(4-27)

318

Chapter 4
The dynamic dynam is the dynam defined by its elements of reduction at o fixed in the
inertial frame:
n

mh a h

h=1

(oa h mh a h )
h =1
o

where a h is the acceleration of the point xh of mass mh ,


or more generally:

a dm
S

r a dm
S
o

where r is the position vector of any point x with acceleration a.


We have:

a dm
d P
.
S
=
dt L o r a dm
G
Go
S
We find again Eq. (4-27) since the dynam of internal forces vanishes.

3.3.3 Conservation of Angular Momentum

PR14 In an inertial frame, the angular momentum of S about a fixed point or the mass center
is constant if the sum of moments of external forces is zero:

M o(e) = 0

Lo = c

(constant).

Proof. It is obvious from the previous result.


This theorem of conservation means that the internal forces of S cannot change the angular
momentum of S, but we recall this is valid if the principle of action and reaction holds.
Remark. The external moment M o(e) must not necessarily be zero so that a first integral
exists; it is sufficient that one of its components vanishes.
Example. Let us consider a system which rotates about the axis oz with an angular velocity .

We have:

Lo = (mh ox h v h ) = (mh ox h ( ox h ))
h

= (mh ox h ( o h x h ))
h

where oh is the projection of xh on the z-axis.

319

Kinetics and Dynamics of Systems

From

Lo = mh [(ox h . o h x h ) (ox h .) o h x h ]
h

and since oh x h is perpendicular to the z-axis, we deduce that the projection of Lo onto this
axis is
( Lo ) z = mh (ox h . oh x h ) = mh o h x h
2

= I z .
We find again a well-known result where I z is the moment of inertia about the z-axis.
If the component of the moment of all external forces along the z-axis (also called the total
external torque about the z-axis) is zero, then the previous remark implies the following
constant:
Iz = c
and two conclusions are possible:
(i) If the system is a rigid body, then the moment of inertia about the axis is constant and the
rigid body necessarily rotates with a constant angular velocity.
(ii) If masses of the system move away from the axis, then I z increases and thus decreases
in the same proportion; on the other hand, if I z decreases then increases in the same
proportion.
Examples illustrate these conclusions, as rotation of skaters, rotation of rotor blades of
helicopters, etc. In first approximation, if we consider that the solar system is isolated from
the rest of the galaxy, then the (total) angular momentum of planets about the mass center of
the solar system is conserved.

3.3.4

Special Case of Rigid Bodies

We consider a rigid body B which is free to translate or rotate with respect to an


inertial frame of reference
= { o; e1 , e 2 , e3 }.

The expression of the angular momentum of B about O is

LO = Ox v x dm
S

where v x denotes the velocity of x with respect to R.


The components of LO relative to the basis (e i ) are

(LO ) i = ipq X p v q dm ,
S

where the components of the velocity v x = v O + Ox relative to (e i ) are


v q = vOq + qrs r X s .
Thus we have

320

Chapter 4

( LO ) i = ipq X p (vOq + qrs r X s ) dm


S

= ipq vOq X p dm + ipq qrs X p r X s dm .


S

By choosing an orthonormal frame of reference R, the first term is obviously:

ipq M X Gp vOq = M (OG v O ) i .


Since

ipq qrs = ir ps is pr ,
the second term is written:

ir ps X p r X s dm is pr X p r X s dm
S

S ( ir X s X

X r X i ) dm

and the inertia tensor I ir appears.


In conclusion, we have obtained:
(LO ) i = ipq M X Gp vOq + I ir r

(4-28)

( LO ) i = M (OG v O ) i + I ir r .

(4-29)

that is:

These general equations relate components of L and with respect to R.


If the rigid body has a fixed point, this is chosen as reference point and thus v O = 0 .
Otherwise, the center of mass G is taken as reference point O and thus OG = 0 .
In both situations, by denoting the reference point by O, we say:
PR15 In a frame of reference R, each component of the angular momentum about a fixed
point of a rigid body or, failing this, about the center of mass is a linear function of all
components of the angular velocity vector:
( LO ) i = I i1 1 + I i 2 2 + I i 3 3 .

(4-30)

Proof. The hypothesis reduces Eq. (4-29) to the following


(LO ) i = I ir r
and thus
( LO ) i e i = I ir r e i

),
i

that is explicitly:
r
1
I 11 I12 I13 I1r
L1

2

r
L2 = I 21 I 22 I 23 = I 2 r .

I
3
L
r
I
I

33
3 O 31 32
3r

Kinetics and Dynamics of Systems

321

We have obtained the essential result:




LO = I .

(4-31)

This relation shows the linear transformation defined by the inertia tensor I and expresses the
angular momentum LO in function of .
We know that vectors LO and have not the same dimensions, more precisely we recall that
the dimensions of the second-order tensor I are ML2 .
The three equations (4-30) refer to axes considered as fixed in space. The inertia tensor is
obviously not constant with reference to these axes, it changes when the body rotates. This
last difficulty may be avoided by choosing a set of axes fixed in the rigid body; then the
corresponding components of I remain evidently constant during the motion of B. This is all
the more helpful since angular momentum theorem makes use of the time derivative of
angular momentum.
The problem is more simplified if these axes are chosen as principal axes since the inertia
tensor is diagonal.
The inertia tensor is written with respect to the frame R E = { O; E1 , E 2 , E 3 } fixed in B and
made up of principal X,Y,Z-axes as following:

I X 0 0
0 I
0
Y

0 0 I Z
where I X , I Y and I Z are constants.

In general use, the angular velocity vector is written with respect to the frame of principal
axes as:
= p(t ) E1 (t ) + q(t ) E 2 (t ) + r (t ) E 3 (t )
and the angular moment about O is expressed as


LO = I X p(t ) E1 (t ) + I Y q(t ) E 2 (t ) + I Z r (t ) E 3 (t ) .

(4-32)

The last two expressions clearly show that and LO are not in general parallel. However,
there are two important special cases where LO and are collinear:
(i) If the three principal moments of inertia are equal (the corresponding ellipsoid of inertia
is a sphere), then:
LO = I

(4-33)

for any direction of .


(ii) If the rigid body rotates about one of the principal axes of inertia (for instance the Z-axis),
then two components of vanish (for instance p and q) and we have necessarily:
LO = I Z

where I Z is the principal moment of inertia about the axis of rotation.

(4-34)

322

Chapter 4

Remark and Example. A principal axis of inertia is sometimes defined as any axis of
rotation parallel to the angular momentum.
Let us illustrate this seeing by considering a system composed of two spheres of mass m
which turn about the axis OZ perpendicular to the straight line joining the centers of spheres,
with angular velocity vector as shown in the first following figure.

Fig. 75
The distance between the mass center c of any sphere and the axis being denoted by R, the
angular momentum due to the rotation of one sphere is
Oc m v ,

that is a vector along the Z-axis (thus collinear to ) and of norm equal to R m R .
The angular momentum of the system about O is thus

LO = 2mR 2 .
This illustrates the previous case (ii) where the axis of rotation is a principal axis of inertia
given the symmetry of the system.
On the other hand, if the angle between the straight line joining the centers of spheres and axis
of rotation is 90 0 , then the principal axis of inertia is not parallel to the axis of rotation.
The angular momentum of a sphere is a vector perpendicular to the straight line joining the
spheres and the norm of which is equal to R m ( R sin ) .
The angular momentum of the system about O is
2 mR 2 sin 1Z .

The system does not rotate about a principal axis of inertia, but in this case LO and are not
collinear.
This concludes the special case of rigid bodies a detailed study of which will be developed in
a next volume.

Kinetics and Dynamics of Systems


3.4

KINETIC ENERGY THEOREMS

3.4.1

Kinetic Energy Theorem

323

3.4.1a Noninertial Frame

Let us consider the motion of a system S of N particles with respect to a noninertial


frame.
PR16 In a noninertial frame R E , the change of kinetic energy due to the motion of S during
a time interval equals the corresponding work done by external, internal and transport
forces exerted on S.
Proof. Given a noninertial frame of reference, we know that the force acting on a particle p h
of S is
d
(mh v h ) = f h(e ) + f h(i ) + f hT + f hC
dt
where we recall that f h( e ) , f h(i ) , f hT and f hC are successively the resultants of external,
internal, transport and Coriolis forces exerted on p h .
From
d 1
( 2 mh v h2 ) = ( f h( e) + f h(i ) + f hT + f hC ) . v h
dt

we deduce the following expression of kinetic energy theorem:




dT
= P (e) + P (i ) + P T
dt

(4-35)

where T is the kinetic energy of S; P ( e) , P (i ) and P T are respectively the external, internal
and transport powers.
No Coriolis power appears since, for any p h , Coriolis acceleration a hC is perpendicular to
velocity v h in any relative frame.
From
N

dT = ( f h( e) + f h(i ) + f hT ) . drh ,
h =1

where drh = v h dt , we deduce:


T T0 =

t0

( f h(e) + f h(i ) + f hT ) . v h dt .

(4-36)

3.4.1b Inertial Frame

Let us consider the motion of a system S of N particles with respect to an inertial


frame R.

324

Chapter 4

PR17 In an inertial frame, the change of kinetic energy due to the motion of S during a time
interval equals the corresponding work done by external and internal forces exerted on
S:
T T0 =

t0

( f h(e) + f h(i ) ) . drh

(4-37)

Proof. It is obvious since, in an inertial frame, the kinetic energy theorem has the following
expression:
dT
= P (e) + P (i ) .
dt

(4-38)

Remark. Even if the system of internal forces is equivalent to zero, the power developed by
these forces does not vanish obviously.
Therefore, the condition of zero external power is not sufficient to obtain a theorem of
conservation of kinetic energy.

We are going to consider the Koenig expression of kinetic energy encountered in Section 2.2.
PR18 The kinetic energy due to the motion of S relative to a frame R 'e = { G; E1 , E 2 , E 3 }
translating with respect to an inertial frame of reference R has a form analogous to the
one in R :
d 1
[ 2 mh (v hR ) 2 ] = ( f h( e ) + f h(i ) ) . v hR .
dt
h
h

(4-39)

Proof. The expression (4-11) of kinetic energy in R leads to the following derivative (with
respect to R):
dT d 1
d
= ( 2 M v G2 ) + 12
[ mh (v hR ) 2 ]
dt dt
dt h

but it is also

= ( f h( e ) + f h(i ) ) . (v G + v hR )
h

and thus

= f h( e) . v G +
h

f h(i) . v G + ( f h(e) + f h(i ) ) . v hR


h

From this last equality, we really deduce Eq. (4-39) since


dv
d 1
( 2 M v G2 ) = M v G G = f h( e ) . v G
dt
dt
h
and since

f h(i) = 0 .
h

[because Eq. (4-17)]

Kinetics and Dynamics of Systems

325

3.4.1c Internal and External Forces Derivable from Potentials

Let

R = { o : e , e , e } be an inertial frame of reference.


1

First, we consider the case where the internal forces are derivable from a potential function
V (i ) , but the external forces are assumed not to be derivable from a potential.
The internal force exerted by the particle of position vector x k on the particle of position
vector xh is denoted
f hk = c hk rhk
where chk R , rhk = x k x h .
The power developed by the internal forces acting on any particle p h is
N

chk rhk . v h .

k =1
k h

Thus, the power developed by all the internal forces is


N

h =1

k =1
k h

P (i ) = ( c hk rhk . v h ) .

(4-40)

By forming pairs {h,k} such that {1,2}, {1,3}, {2.3} and so on, the previous power is written:
N

h =1

k =1
h<k

P (i ) = ( c hk rhk . (v h v k ))
N

h =1

(4-41a)

d
( c hk rhk rhk )
dt
k =1
h< k

also written:

P (i ) =
The factor

1
2

chk rhk . r&hk

2 h ,k

(4-41b)

is necessary since by summing over k and h, each point of pair is twice

considered.
PR19 In an inertial frame, if the internal forces of a system of N particles derive from an
(internal) potential V (i ) , then the time derivative of the sum of kinetic energy and
internal potential equals the power of external forces:
N
d
(i )
(T + V ) = f h( e) . v h .
dt
h =1

Proof. If we introduce the following function internal potential defined by

(4-42)

326

Chapter 4
N

V (i ) ( x1 ,..., x N ) =

chk rhk . drhk

h =1 k =1
h<k

we have:
N
V (i )
=
rhk
h =1
h< k

chk rhk
k =1

and thus
N
dV (i )
=
dt
h =1
h< k

chk rhk dt rhk


k =1

= P (i )

and the kinetic energy theorem (4-38) leads to the result.


Now, in addition if the external forces are derivable from a potential energy V (e ) , that is

f h( e ) = hV (e )

h =1,..., N

then the theorem of kinetic energy is expressed as follows:


PR20 In an inertial frame, if all internal and external forces of a system S derive from a
potential then the mechanical energy is constant:
T +V = c .

(4-43)

Proof. Since

f h( e ) . v h = (

V ( e)
V ( e)
V (e )
+
e
e
+
e3 ) . v h
2
1
x1h
x h2
x h3

the theorem of kinetic energy and more precisely Eq. (4-42) becomes:
d
(T + V (i ) + V (e ) ) = 0
dt

which proves the proposition.


Remark. In general, there are internal frictional forces (which depend upon relative
velocities of particles) and the conservation theorem no longer holds.

3.4.2

Special Case of Rigid Bodies

3.4.2c Expression of Kinetic Energy

First let us consider a rigid body B which rotates about a given axis .
Let = & be the angular velocity of B about and let Oh denote the projection of any point
x h B of mass mh .

327

Kinetics and Dynamics of Systems

We know that the kinetic energy of B is

T=

1
2

mh (

O h x h & ) 2 =

1
2

I 2

where I is the moment of inertia about .


This expression is not very interesting since the axis position is generally unknown.
In a general manner, we consider a rigid body B which is free to translate or rotate with
respect to an inertial frame of reference
= { o; e1 , e 2 , e3 }.

The kinetic energy of B is

T=

1
2 S

v 2 ( x) dm =

1
2 S

v i v i dm

).
i

Given O B , the components relative to (e i ) of the velocity of any point x B are written:
v i = vOi + ijk j X k
and thus
v i v i = vOi vOi + 2 ijk vOi j X k + ijk irs j X k r X s .
But

ijk irs = jr ks js kr
and thus
T = 12 M vOi vOi + ijk vOi j X k dm + 12 ( jr ks js kr ) j r X k X s dm
S

1
2

M vOi vOi

M ijk vOi j X Gk

1
j r jr
2

S X s X

dm

1
j r
2

S X r X j dm

that is
T = 12 M vOi vOi + M ijk vOi j X Gk +

1
2

I jr j r .

(4-44)

In conclusion, we have obtained:


T = 12 M v O + M v O . OG +
2

1
2

I jr j r .

(4-45)

If the rigid body has a fixed point, this is chosen as origin of frame of reference and we say:
PR21 Given a frame of reference R such that the origin is the fixed point of a rigid body B,
the kinetic energy is only due to the rotation of B about the fixed point:
T = 12 I ij i j .

(4-46)

Otherwise, the center of mass of B is to be taken as reference point O B , thus OG = 0 , and


we say:

328

Chapter 4

PR22 Given a frame of reference R, the kinetic energy of B is the sum of the translational
kinetic energy associated with the motion of G and the rotational kinetic energy
associated with the rotation of B about G:
2

T = 12 M v G +

1
2

I ij i j .

(4-47)

We note that the quadratic form 12 I ij i j represents the kinetic energy due to the rotation.
This intrinsic scalar being independent of the choice of reference axes, we can choose the
(variable) axis of rotation as Z-axis such that X = Y = 0, Z = . We find again the wellknown:
T = 12 I Z 2 .

Components relative to axes fixed in the rigid body will not change with time. In addition,
we know that the most convenient axes are the principal axes of the inertia tensor (also called
the principal axes of the rigid body B).
In this case where the principal moments of inertia are constants denoted I X , I Y and I Z , the
kinetic energy is explicitly written:
T = 12 ( I X X2 + I Y Y2 + I Z Z2 )

T = 12 ( I X p 2 + I Y q 2 + I Z r 2 ) .

(4-48)

3.4.2b Theorem of Kinetic Energy

PR23 The time derivative of the kinetic energy of a rigid body B equals the power of
external forces exerted on B:
dT
= P (e)
dt

(4-49)

that is
f (e)

(e) .
M O v O

Proof. First, we know the power of internal forces is zero.


Indeed, we recall
P (i ) = f h(i ) . v h = f h(i ) . (v O + Ox h )
h

= f

(i )

. v O + . (Ox h f h(i ) ) ,
h

but we know that the dynam of internal forces of any rigid body is zero:
f (i )
(i ) = 0 .
M O

329

Kinetics and Dynamics of Systems

Since
P

(i )

f (i )
=

(i )
M O

v = 0 ,
O

the kinetic energy theorem is really


dT
= P (e)
dt

where
P

( e)

f (e)
=

(e)
M O

v .
O

PR24 The linear and angular momenta are expressed from the kinetic energy as follows:
(P ) i =

T
vOi

(LO ) j =

T
j

(4-50)

and the kinetic energy is also written:


P
T = 12
LO

.
vO

(4-51)

Proof. From (4-44), we deduce:


T
= M [(v O ) i + ijk j X Gk ]
i
vO
= ( P )i
(see Example of Section 3.2.4).
From (4-28), we deduce:
T
= M ( ijk vOi X Gk + I jr r )
j

= M ( jki X Gk vOi + I jr r )
= ( LO ) j .
Moreover, given the homogeneous quadratic function T, Eulers theorem leads to the
following:
T
T
vOi
= 2T
+ j
i
j
vO
that is
2 T = v O . P + . LO

and the proposition is proved.


In particular, we may state:

330

Chapter 4

PR25 If a rigid body rotates about a fixed point or if its motion is considered about its mass
center, then its kinetic energy is
T = 12 . LO .

(4-52)

Example. Two homogeneous slender bars, each of mass m and length l, are pinned together.
Express the kinetic energy of the system of bars moving in a vertical plane in function of the
angle as shown in Fig. 76.

Fig. 76
The kinetic energy of the bar oa is
l
I
Toa = 12 m ( & ) 2 + & 2 .
2
2

Since & is the angular velocity of each bar, we have:


l
I
Tab = 12 m [( 32 l ( sin )& ) 2 + ( & cos ) 2 ] + & 2 .
2
2

Since the moment of inertia I about the axis of rotation through the mass center of respective
bars is m l 2 12 , we obtain:

T=

ml 2 & 2 ml 2
ml 2 & 2
+
(9 8 cos 2 )& 2 +

&2

12

= ml ( 43 cos ) .
2

331

Kinetics and Dynamics of Systems

4. EXERCISES
Exercise 1.
A vertical homogeneous plate P of mass M has an isosceles triangle shape. This
triangle oac is such that oc is vertical, oc = 2l and oa = ac = 2 l . A horizontal
homogeneous disk D of mass M , of center o and radius R is welded on to the plate at o.

(i) Find the inertia tensor of P, at the point h which is the orthogonal projection of a on oc.
(ii) Determine the position of the center of mass G of the system S made up of P and D.
(iii) Find the inertia tensor of S at point o.
(iv) Calculate the kinetic dynam and the dynam of forces when the system S rotates about oc
with the angular velocity & . Express the kinetic energy of S.
Answer. (i) Let R e = { o;1x ,1 y ,1z } be the fixed frame of reference, which at point h is
written { h;1x ,1 y ,1z }. Let R E = { o;1X ,1Y ,1Z } be the moving frame fixed in S such that

1X (t ) is collinear to ha and 1Z (t ) is collinear to hc.


The material symmetries imply:
I X = (Y 2 + Z 2 ) dm = Z 2 dm ,
P

I Z = ( X 2 + Y 2 ) dm = X 2 dm
P

and thus
IY = I X + I Z .

Fig. 77
We have:
l

I X = dX
0

M l2
6

lX

X l

Z 2 dZ =

((l X )
3 0

( X l ) 3 ) dX =

l4

332

Chapter 4
l

I Z = X dX
2

l X

X l dZ =

M l2
6

So, the inertia tensor of P at point h is

I P( h )

Ml2
0

Ml2
= 0
3

0
0

0 .

2
Ml

6
0

(ii) From
( M + M ) oG = M oG1 + M 0 ,
we deduce:
oG =

Ml
M
oG1 =
( 1Z + 13 1 X ) ,
M + M
M +M

where G1 is the center of mass of P.


(iii) The inertia tensor of S, at o, is obviously the following sum:

IS(o ) = I P(o ) + I D( o ) .
The inertia tensor of the disk D, at o, is well-known:

I D(o )

M R 2

0
0

M R 2
= 0
0 .
4

M R
0
0

We are going to calculate IP(o ) by using the Steiners theorem twice.


We have:
I (pqo ) = I (pqG1 ) + M ( pq d 2 x (pG1 ) x q(G1 )
where d = oG1 .
In addition, from
I (pqh ) = I (pqG1 ) + M ( pq d 2 x p(G1 ) x q (G1 ) )
where d' = hG1 , we deduce:
I (pqG1 ) = I (pqh ) M ( pq d 2 x p(G1 ) x q (G1 ) ) .

333

Kinetics and Dynamics of Systems

Therefore, we have obtained the following result:


I (pqo ) = I (pqh ) M ( pq d 2 x p(G1 ) x q (G1 ) ) + M ( pq d 2 x p x q ) .
So, in particular, we have successively:
(o)
I11
=M

7
Ml2,
6

(o)
I 22
=M

l2
l2
M ( y 2 + z 2 ) + M ( y 2 + z 2 ) = M
0 + Ml2
6
6

4
3

l2

M ( x 2 + z 2 ) + M ( x 2 + z 2 ) = M

l2
3

l2
9

+M

10 l 2
9

Ml2,

(o)
I 33
=M

=M

l2

M ( x 2 + y 2 ) + M ( x 2 + y 2 ) = M

l2
6

l2
9

+M

l2
9

(o)
I12
= 0 M ( x1 x 2 ) + M ( x1 x 2 ) = 0 ,
(o)
I13
= M

l2
3

( 0)
I 23
= 0.

Thus, the inertia tensor of P at o is

I P(o )

7 M l 2
0 13 M l 2
6

4 Ml2
.
= 0
0
3

1 Ml2
13 M l 2 0
6

In conclusion, the inertia tensor of S at o is

IS(o )

1 M R 2 + 7 M l 2

13 M l 2
0
6
4

2
2
1
4

=
0
0
4 M R + 3Ml

2
1
1 M R 2 + 1 M l 2

M
l
0
3
2
6

(iv) The linear momentum of S with respect to R e is


P = M v G1 + M v o = M (v o + G1o S R e ) = M& 1Z l ( 13 1X + 1Z )
= 13 M l& 1Y .

Since o is fixed in R e , the angular momentum of S about o with respect to R e is

334

Chapter 4
3

Lo = I is s E i (t )
i =1

where ( E1 , E 2 , E 3 ) = (1X ,1Y ,1Z ) .


We have:
3

i =1

i =1

Lo = I i 3 3 E i = I i 3 & E i
=

Ml2 &
M l 2 M R 2 &
E1 + (
+
) E 3 .
3
6
2

The kinetic dynam of S with respect to R e is expressed, at o, as

M l&
E2

.
M l 2&
M l 2 M R 2 &
+
E1 + (
) E 3

3
6
2

o
Let us calculate the dynamic dynam.
We have:
deP Ml d E &
=
( ( E 2 ) + Ee & E 2 )
dt
3 dt
M l &&
=
( E 2 & 2 E1 ) ,
3

since
E1

E2

Ee & E 2 = 0

E3
& = & 2 E1 ;

0
&

and we also have:


M l 2 M R 2 &
M l 2&
M l 2 M R 2 &
de
d E M l 2&
[
) E 3 ] + Ee [
) E 3 ]
Lo =
E1 + (
E1 + (
+
+
6
3
6
2
3
2
dt
dt
M l 2&&
M l 2 M R 2 &&
M l 2&
&
) E 3 + E 3 (
E1 + (
E1 )
=
+
3

Ml
3

(&& E1 + & 2 E 2 ) + (

Ml
6

M R
2

)&& E 3 .

In conclusion, the dynamic dynam of S with respect to R e is


M &&

&2
3 ( E 2 E1 )

.
M M R 2 &&
M &&
2
&
3 ( E1 + E 2 ) + ( 6 + 2 ) E 3
o

Kinetics and Dynamics of Systems

335

In this problem, the kinetic energy T = 12 I ij i j of S is written as it follows


76 M l 2 + 14 M R 2
0
0
13 M l 2


2
2

0 ,
4
1
1
&

0
0
T = 2 (0 0 )
3Ml + 4M R


2
2 &
1
1
1M l2

0
3
6 M l + 2 M R

that is
T = ( 121 M l 2 + 14 M R 2 )& 2 .

Exercise 2.
The ends a and b of a homogeneous rectilinear rod B of mass M and length R 3 move
along a hoop C of center o and radius R. The hoop rotates about its vertical fixed diameter.
In a frame of reference R e = { o; e1 , e 2 , e3 } such that e3 is along this diameter:
(i) Express the kinetic and dynamic dynams of B at o,
(ii) Find the kinetic energy of B.
Answer.

Fig. 78
(i) By considering the unit vector E1 orthogonal to the plane of the hoop C, the moving
frame R C = { o; E1 , E 2 , E 3 } fixed in C is obtained from R e by a rotation of angle about
the vertical diameter.
In the plane of C, the rotation of angle about E1 is the following transformation:
{ o; E1 , u, e 3 } { o; E1 , E 2 , E 3 }
where E 2 is along the rod B and u is the horizontal unit vector in the vertical plane of C.
The linear momentum of B is expressed in ( E1 , E 2 , E 3 ) as follows:

336

Chapter 4
d e oG
de R
=M
( E3 ) ,
P=M
dt
dt 2

where G is the mass center of the rod B,


but
de
dE
E3 =
E 3 + Ee E 3 = (& e 3 + & E1 ) E 3
dt
dt

and thus
P=

M
2

R (& sin E1 & E 2 ) .

Let us calculate the angular momentum of B:


3

Lo = I is s E i .
i =1

The inertia tensor of B with respect to ( E1 , E 2 , E 3 ) , at point G, is the well-know tensor:

I (G )

M 2
R
4
= 0

0
0
0

0 .

M 2
R
4

From
I (pqo ) = I (pqG ) + M ( pq d 2 x (pG ) x q(G ) )
where
M ( yG2 + z G2 ) =

M
4

R2 ,

M ( xG2 + z G2 ) =

M
4

R2 ,

we deduce that

I (o )

Since
we deduce that

M 2
0
0
R

M 2
R
0 .
= 0
4

M 2
0
R
0

= & e 3 + & E1 = & E1 + & sin E 2 + & cos E 3 ,

M ( xG2 + y G2 ) = 0 ,

337

Kinetics and Dynamics of Systems


M 2

0
0
R
2
&

M 2
Lo = 0
0 & sin E i
R
4
i =1
& cos
M

0
0
R 2

=
=

M
2
M
2

M
M
R 2 & E1 + R 2& sin E 2 + R 2 & cos E 3

R 2& E1 +

4
M
4

R 2& e3 .

The elements of the dynamic dynam, at o, are obtained as follows.


First, we have:
deP M R de &
( sin E1 & E 2 )
=
dt
2 dt
d e E1 &&
d e E2
M R &&
[( sin + & & cos ) E1 + & sin
]
=
E 2 &
2
dt
dt

but
d e E1
= E1 = & (e 3 E1 ) = & u ,
dt

d e E2
= E 2 = (& e 3 + & E1 ) E 2
dt
= & sin(90 o ) E1 + & E 3
and thus:
d e P M R &&
=
( sin + 2&& cos ) E1 + & 2 sin u && E 2 & 2 E 3 ) .
dt
2

Secondly, we have:
d e Lo M 2 &&
M 2&&
M 2 &&
R E1 +
R u+
R e3 .
=
dt
2
2
4

(ii) The kinetic energy of the rod B is


M
M
T = 12 (. Lo ) = 12 (& e 3 + & E1 ) . ( R 2& E1 + R 2& e 3 )
2

M
4

R 2& 2 +

M
8

R 2& 2 .

CHAPTER 5

LAGRANGIAN DYNAMICS
VARIATIONAL PRINCIPLES

Lagrangian dynamics is not a new mechanics with respect to Newtonian mechanics,


but it is another formulation of motion equations from the notion of virtual displacements.
Instead of considering the notion of force, the Lagranges method introduces the kinetic
energy (and also potential energy and Lagrangian). The interest of this method is it uses scalar
functions, which are invariant under coordinate transformations!
Passing from a space in which the vector equations of motion may not be obvious to a
corresponding space of generalized coordinates is often a simplification. We recall in
particular that the forces of constraint are generally unknown while the mechanical energy of
a system can be easily known.
Hamiltons variational principle states that nature acts in a determined manner: to
minimize the time integral of the difference between kinetic and potential energies.
Lagranges equations have a very simple form and are easily obtained. Only a
preliminary difficulty consists in the more judicious choice of generalized coordinates.

1. LAGRANGIAN DYNAMICS
Let E be a 3-dimensional Euclidean vector space,
E be the Euclidean point space associated with E,
{ o; e1 , e 2 , e3 } be a frame of reference,
be an open of E,
I be an open of R,
U be an open of R n .

339

340

1.1

Chapter 5
HOLONOMIC AND SCLERONOMIC SYSTEMS

Let us consider a system S of N particles with a finite number n of degrees of freedom


and with possible constraints. Briefly, any particle ph is a point xh with a mass mh .
We introduce n functions of class C 2 , called the generalized coordinates:
q i : J ( I ) R : t a q i (t ) .

We will indistinctly denote1:


q = (q 1 ,..., q n ) U .

Each particle p h is located by its position vector and for each p h we consider a vector
function of class C 2 :
rh : I U : (t , q ) a rh (t , q ) .
So, every motion of each particle p h is defined by the position vector

x h = rh (t , q1 ,..., q n )
= rh (t , q).

(5-1)

Notation. According to usage, rh will be identified with x h .


D

A system of N particles p h is holonomic if the position vectors rh depend on


generalized coordinates q i and time t only.
If position vectors depend on q& i (called the generalized velocities), then the system is
said to be nonholonomic.

We recall that a constraint is called holonomic if no time derivative of position vector appears
in any constraint equation:
f (t , r1 ,..., rN ) = 0 .

(5-2)

Otherwise, the constraint is said to be nonholonomic.


A material system with nonholonomic constraints is obviously nonholonomic.
Unless otherwise specified, holonomic systems will be studied.
D

A system of N particles p h is scleronomic if the position vectors rh do not explicitly


depend on time:
rh
= 0.
t

Or simply

q = (q i ) where it is an understood thing that i {1,,n}.

341

Lagrangian Dynamics, Variational Principle

We recall that a constraint is called scleronomic if it is not explicitly dependent on time:


f
= 0.
t

Otherwise, the constraint is said to be rheonomic.


A material system with rheonomic constraint is obviously rheonomic.
Example. In a vertical plane defined by { o; e1 , e 2 } with e 2 upwards, we consider a simple
pendulum consisting of a particle p suspended by a weightless rigid rod of length l which
oscillates from an extremity O according to

oO = sin 2t e1 .
The position vector of p is
r (t , ) = (sin 2t + l sin ) e1 l cos e 2
where is the angle (e 2 , Op) .
In general, since the position vector of any particle of a rheonomic system is
x h = rh (t , q(t )) ,
we obtain the following expressions (with summation over i and j):
x& h =
x&&h =

rh
r
(t , q) + hj (t , q) q& j ,
t
q
2 rh
t

(t , q ) + 2

2 rh
tq

(5-3)

(t , q ) q& +
j

rh
q

(t , q ) q&& +
j

2 rh
q q
i

(t , q ) q& i q& j .

(5-4)

Remark. A rheonomic system can present a Lagrangian that is not explicitly dependent on
time.

Indeed, let us consider a pearl following a circle c(o;R). If this circle rotates about a vertical
axis with a constant angular velocity , then this example illustrates this remark.
The generalized coordinate is the angle (in the plane of the circle) locating the pearl from
the horizontal plane of equation x 3 = 0 .
The position of the circle is located by the angle t in the horizontal plane and is so
explicitly dependent on time.
The components of the position vector of the pearl are:
x1 = R cos cos t ,

x 2 = R cos sin t ,

The Lagrangian
L=

m
2

( R 2& 2 + R 2 2 cos 2 ) mg R sin

is not explicitly dependent on t.

x 3 = R sin .

342

1.2

Chapter 5
DALEMBERT-LAGRANGE PRINCIPLE

The idea of dAlembert was to make dynamics similar to statics and to use the statics
methods in dynamics (1743).
More precisely, this principle consists in considering the equations of motion of every particle
ph :
f h mh a h = 0
as a condition of dynamic equilibrium.
The consideration of inertia forces mh a h creates an artificial state of equilibrium and the
dAlemberts principle states:
PR1

A system S of particles is in equilibrium under the action of all applied forces and
inertia forces.

It is expressed as
p h S : f h mh a h = 0

(5-5)

where f h represents all the applied forces acting on p h , that is the resultant of given
(external and internal) forces Fh and unknown forces Lh of constraint on p h :
f h = Fh + Lh .
Later, Lagrange used the concept of virtual displacement.
We recall that a virtual displacement of a point xh is a priori an arbitrary vector, denoted by
x h (time being fixed: t = 0 ).
From above, we have obviously:
N

(Fh + Lh mh a h ). x h = 0 .
h =1

The dAlembert-Lagrange principle consists in choosing virtual displacements for which the
(unknown) forces of constraint do not virtually work.
Such virtual displacements, said to be compatible with the constraints, must verify the
following:
N

Lh . x h = 0 .

(5-6)

h =1

So, in the context of virtual work, dAlembert-Lagrange principle is expressed as follows:




(Fh mh a h ) . x h = 0 ,

(5-7)

h =1

that leads to
PR2 The motion of a system S is such that the virtual work of the resultant of given applied
forces and inertia forces is zero.

Lagrangian Dynamics, Variational Principle

343

Remark. We note that the various x h are generally not independent and it is out of
question to vanish the individual terms of the previous sum (unlike f h mh a h = 0 ).
Example. Let us find again the equation of motion of a rigid body which rotates about a
fixed axis .

The system has one degree of freedom and the angle of rotation of the body is chosen as
generalized coordinate .

Fig. 79
A virtual displacement of any point xh of the body is

x h = ox h
where o is an arbitrary point of .
The dAlembert-Lagrange principle leads to

(Fh(e) mh a h ) . x h = 0 .
h

First, we have

Fh(e) . x h = Fh(e) . ox h = . M o(e)


h

where M o(e) denotes the moment about o of the given external forces.
Second, we consider

m h a h . x h
h

and we note that the normal acceleration of xh does not

contribute to this virtual work. Since the tangential acceleration has a norm equal to oh x h &&
and since
x h = ox h = oh x h = oh x h ,
we have:

344

Chapter 5

m h a h . x h = m h o h x h
h

&& .

Therefore, the equation following from the dAlembert-Lagrange principle is


I && = . M o(e ) = 1 . M o( e)
= [ M o(e ) ]

where [ M o(e ) ] is the (algebraic) projection M o(e ) . 1 of M o(e) on .

We have obtained the well-known equation of motion:


I && = [ M o(e) ] .

1.3

LAGRANGES EQUATIONS

We are going to obtain the famous Lagranges equations from the dAlembertLagrange principle [Eq. (5-7)].1
1.3.1

Lagranges Equations in the General Case

We deal with the first term of Eq. (5-7).


Given the position vector x h = rh (t , q) of any particle p h , we know that

x h = ( x h ) i e i =

rh
q

qj

( j = 1,..., n ).

We also know that the generalized force associated with the jth generalized coordinate and
relating to the given forces Fh is the function Q j (t , q, q& ) defined by
N

rh

h =1

q j

Q j = Fh

(5-8)

where the terms of the scalar product are explicitly:


Fh (t , rh (t , q), r&h (t , q, q& ))
and
rh
q j

(t , q) .

The virtual work W = Fh . x h done by the forces Fh during virtual displacements x h


h

is immediately:
1

Of course this presentation is different from the Lagranges one made in his remarkable treatise Mcanique
Analytique (1788).

Lagrangian Dynamics, Variational Principle

Fh . qhj q j = Q j q j .

345

(5-9)

h =1

In regard to the second term of Eq. (5-7), we are going to express

mh a h . x h

in function

of the kinetic energy of the system.


First, we express:
dx& h
dr& r
x h = mh h hj q j
dt
dt q
h
r
d
d rh
mh ( (r&h hj ) r&h
) q j ,
j
dt
dt
q
q

m h a h . x h = m h
h

=
h

(5-10)

where we have identified x h with rh according to usage.


Intermediate remark. We must clarify the following point which temporarily consists in
giving up the sense attached to quantities.

For example, we will not consider the velocities x& h expressed as composite functions of t any
more, but as functions of 2n + 1 independent variables t , q j , q& j .
To obtain the Lagranges equations, we emphasize that we are going to consider the 2n + 1
variables as being independent a priori, that is for instance:

q& j = 0 .

We consider rh (t , q ) and not rh (t , q(t )) any more!


Since t and various q j are independent, we deduce:

drh =

rh
r
dt + hj dq j
t
q

with independent t , dt , q j , dq j and so we have:

drh rh rh j
=
+
q&
t q j
dt
where t , q j , q& j are independent variables.
The following calculations are made by considering all the various q j , q& j ,... as independent
variables.
Only later does q (t ) take the place of q where t plays a special role; the successive
&&h (t ),... are (generally)
q& (t ), q&&(t ),... are (generally) calculable from q (t ) . The various x& h (t ), x
calculable from x h (t ) = rh (t , q (t )) .
But the successive derivatives

346

Chapter 5

x& h (t ) =

rh
r
(t , q(t )) + hj (t , q(t )) q& j (t ) ,
t
q

&x&h (t ), &x&&h (t ),

become identities (equalities for every t). This concludes this essential remark.
Now, we come again to Eq. (5-10).
From
r
r
r&h (t , q, q& ) = h (t , q) + hj (t , q) q& j
t
q
we deduce:
r&h rh
=

q& i q i

(1)

In addition, we have:
2 rh
2 rh
d rh
(t , q) =
+ j i q& j
i
i
dt q
q q
tq
r
r
= i ( h + hj q& j ) ,
q t q
since rh (t , q ) being of class C 2 , then the partial derivatives are permutable.
Thus, we have obtained:

d rh r&h

=
dt q i q i

(2)

By taking (1) and (2) into account, Eq (5-10) is written:


r&

r&

mh a h . x h = mh [ dt (r&h q&hi ) r&h qhi ] q i


h

= mh [
h

1
d 1
2
2
( r&h ) i ( r&h )] q i .
i
dt q& 2
q 2

By introducing the kinetic energy of the system


T=

mh
2

r&h ,

we have thus obtained:


d T

mh a h . x h = ( dt q& i q i ) q i .

(5-11)

By taking account of Eqs. (5-9) and (5-11), we conclude that Eq. (5-7) expressing the
dAlembert-Lagrange principle is written:
(Qi

d T T
+ i ) qi = 0
i
dt q&
q

347

Lagrangian Dynamics, Variational Principle

and since the n generalized coordinates are independent, we deduce the n Lagranges
equations


d T T
= Qi .

dt q& i q i

(5-12)

Remark 1. We emphasize that first we will calculate the derivatives

T
q

T
where the
q& i

various q i and q& i are considered as independent variables. Next, in the so-obtained
expressions, we will replace the various q i by q i (t ) and q& i by

dq i
(t ) .
dt

Remark 2. The previous Lagranges method leads to the determination of motion of a


material system while the forces of constraint are ignored, which is interesting since these are
generally unknown.
However, we note that the fundamental law
f h mh a h = 0
leads to

( f h mh a h ) . rh = 0
h

for every virtual displacement rh and where terms f h also take constraint forces into
account.
So, if we denote the generalized force associated with q j and relating to the resultant of all
forces (including the forces of constraint) by
N

rh

h =1

q j

Q j = fh

(5-13)

then we have:
N

f h . rh = Q j q j .

h =1

But since
d T

mh a h . rh = ( dt q& j q j ) q j ,
h

we deduce:
d T
T
j = Qj ,
j
dt q&
q

(5-14)

which shows that Lagranges equations are very general and can take the constraint forces
(within Q j ) into account.
Of course, the interest of the Lagrangian method is to choose virtual displacements which
vanish the contribution of unknown forces of constraint to the virtual work.

348

Chapter 5

Classical example. Find again the equations of motion of a free particle of mass m in
spherical coordinates relative to an inertial frame { o; i , j , k }.
Given the three generalized coordinates
q2 = ,

q1 = r ,

q3 = ,

the Cartesian coordinates are expressed as


x = r sin cos ,

y = r sin sin ,

z = r cos .

The generalized forces associated with the various generalized coordinates q j and relating to
the applied force F of components Fx , F y , Fz are successively:

(x i + y j + zk)
r
= Fx sin cos + Fy sin sin + Fz cos ,

Qr = F

Q = r ( Fx cos cos + Fy cos sin Fz sin ) ,


Q = r sin ( Fy cos Fx sin ) .

From
T=

m
2

(r 2& 2 + r 2& 2 sin 2 + r& 2 )

we deduce the following motion equations of the free particle:


d
(mr&) mr (& 2 + & 2 sin 2 ) = Qr ,
dt
d
m
(mr 2& ) r 2& 2 sin 2 = Q ,
2
dt
d
(mr 2& sin 2 ) = Q .
dt

1.3.2 Lagranges Equations for Conservative Forces


If all (internal and external) given forces are derivable from a potential V, then the
corresponding virtual work is written:
Q j q j = Fh . rh = hV
h

rh
q

qi

V
qi .
i
q

Since the generalized coordinates are independent, we deduce that the generalized forces are
Qj =

V
q j

(5-15)

Lagrangian Dynamics, Variational Principles

349

Lagranges equations are written:


d T T V
=0
+

dt q i q i q i

and since the potential does not depend on various q j , these equations are also:
d

(T V ) i (T V ) = 0 .
i
dt q
q

Let us introduce the following


D

The Lagrangian of a material system is the function


L : I U R n R : (t , q, q )  L(t , q, q )

such that
L(t , q, q ) = T (t , q, q ) V (t , q ).

(5-16)

Therefore, Lagranges equations for systems subject to conservative forces are simply:
d L L

=0.
dt q i q i

(5-17)

They are n in number if the material system has n degrees of freedom.


PR3

Lagranges equations are invariable under changes of generalized coordinates.

Proof. Given any change of generalized coordinates denoted by


q j = q j (t , q )

j = 1,..., n

and if we denote:

L (t , q , q ) = L(t , q(t , q ), q (t , q , q )) ,
then we have:
L q j
L
L q j

+
=
q i q j q i q j q i

Since
q j
=0
q i

and
q j
q j
dq j
q j q j k

,
+
q
=
=
(
)
=
(
)
q k
q i
q i
q i dt
q i t

we have:

350

Chapter 5
L
L q j
=
q i q j q i

which implies:
d L
d L q j
L d q j

+
=
dt q i dt q j q i q j dt q i

Besides, we have:
L q j
L q j

+
=
q i q j q i q j q i
L

Finally, we deduce:
0=

d L L
d L
L d q j q j
L q j

=
(
)
(
)
dt q i q i
dt q j q j q i q j dt q i
q i

The last term vanishes because we have:


2 q j
2 q j k d q j
q j q j k

.
+ k q ) = i + i k q =
= i(
dt q i
t
q
q t q q
q i
q

q j

We conclude that
d L
L
=0

j
dt q
q j

since the Jacobian of the transformation

(q )
is different from zero.
(q )

1.3.3 Lagranges Equations with Undetermined Multipliers

We are going to obtain the Lagranges equations given s (overabundant) primitive


coordinates u i and p equations of constraint which are
a ri (t , u 1 ,..., u s ) du i + br (t , u 1 ,..., u s ) dt = 0

(i = 1,..., s )

if the constraints are nonholonomic.


For holonomic constraints, this system of equations is completely integrable, that is there are
p functions r such that:

r (t , u 1 ,..., u s ) = c r
or p equations

(constants),

351

Lagrangian Dynamics, Variational Principles

d r )

r i
r
dt +
du = 0
t
u i

which are so expressed as in the nonholonomic case.


The equations of constraint lead to the p following equations for virtual displacements:
a ri (t , u 1 ,..., u s ) u i = 0

(5-18)

(since t is fixed),
or p following equations for virtual velocities:
a ri (t , u 1 ,..., u s ) u i = 0 .

(5-18)

The dAlembert-Lagrange principle, written for virtual displacements compatible with the
constraints, introduces the following equation
N

rh

h =1

(Qi mh a h

(Qi

) ui = 0

d T T
+ i ) u i = 0
i
dt u
u

(i = 1,..., s )

(1)

where the arbitrary increments u i are no more independent! So, at this stage, we cannot
deduce s equations
d T T
= Qi .

dt u i u i

But this difficulty is overcome by introducing Lagrange multipliers. This Lagranges method
consists in multiplying the p relations of constraints by p successive multipliers (which are
arbitrary parameters a priori) and next in summing; namely:
p

r ari u i = 0 ,
r =1

where the p multipliers r are dependent on time.


Equation (1) is equivalent to the following
(Qi

p
d T T
+
+
r ari ) u i = 0 .
dt u i u i r =1

(2)

From p equations

a ri u i = 0

(i = 1,..., s ) ,

we can deduce p quantities u i , for instance u 1 ,..., u p in function of the ( s p )


others: u p +1 ,..., u s which are independent!
Moreover, by choosing the p multipliers judiciously, we can vanish p terms of equation (2),
for example the first p following

352

Chapter 5
p
d T T
Qi
+
+ r a ri = 0
dt u i u i r =1

i = 1,..., p.

(3)

Only s p coefficients of u i remain in (2) for which u p +1 ,..., u s are independent;


therefore these s p coefficients are necessarily zero; namely:
Qi

p
d T
T
+
+
r a ri = 0

dt u i u i r =1

i = p + 1,..., s. (4)

By putting (3) and (4) together, we have a system of s equations


p
d T T

Q
=
r ari
i
dt u i u i
r =1

i = 1,..., s,

(5-19)

r = 1,..., p.

(5-20)

to which we add p equations of constraint


a ri u i = 0

In conclusion, the problem consists in solving a system of s + p equations with s + p


unknowns.
If the forces are derivable from a potential, Eqs. (5-19) are written:
p
d L L

= r a ri
dt u i u i r =1

i = 1,..., s.

(5-21)

Physical interpretation of Lagrange multipliers

PR4

Each generalized force of constraint is the second member of the corresponding


Lagranges equation with multipliers.

Proof. We must prove that


N

h =1

Lh

rh
u i

= r a ri .
r =1

The dAlembert-Lagrange principle

(mh a h Fh Lh ) . rh = 0
h

leads to the following equations


(

rh
d T T

L
Q
) ui = 0 .

i
h
i
i
i
dt u
u
u
h

By identifying this equation with


(

p
d T T

r a ri ) u i = 0 ,

i
i
i
dt u
u
r =1

(5-22)

Lagrangian Dynamics, Variational Principles

353

we obtain Eq.(5-22).
Example. Find the Lagranges equations with one multiplier of a particle p of mass m
frictionless moving on a cylinder of equation

x2 + z2 = R2.
So we will consider three primitive coordinates x, y, z of this particle having two degrees of
freedom and we will denote the position vector of p by
r = x i + y j + zk.

Answer. From the obvious expressions


x = R cos ,

y = y,

z = R sin

we deduce:
L=

m
2

( x 2 + y 2 + z 2 ) mg z.

The equation of constraint x 2 + z 2 R 2 = 0 implies that every virtual displacement


compatible with the constraint must verify:
2x x + 2z z = 0

(in this case: a11 = 2 x, a12 = 0, a13 = 2 z ).


The Lagranges equations with multiplier are
d L L

= 2x ,
dt x x

d L L

= 0,
dt y y
d L L

= 2z ,
dt z z

that is
mx = 2 x ,
my = 0 ,
mz = mg + 2 z .

The Lagrange multiplier is connected with the force of constraint as follows:


L

r
= Lx = 2 x ,
x

r
= Ly = 0
y

354

Chapter 5
L

r
= Lz = 2 z
z

and thus the force of constraint is


L = 2 ( x i + z k ) = 2 R 1n
where
1n = (cos i + sin k )
is the unit vector toward the cylinder axis.
We also obtain the expression of L from the previous Lagranges equations as follows:
L + mg = ma
= mx i + my j + mz k
= 2 R cos i + 2 R sin k mg k
= 2 R 1n mg k .

We note that the equations of motion of p are obviously obtained by considering the
generalized coordinates and y.
Indeed, the Lagrangian being
L = T V =

( R 2 2 + y 2 ) mgR sin ,

these equations are immediately:


d L L
=0,

dt 

d L L

=0
dt y y

1.4

R + g cos = 0 ,

my = 0.

CONFIGURATION SPACE AND LAGRANGES EQUATIONS

A moving mechanical system S made up of particles and of (possibly) rigid bodies is


assumed to have a finite number n of degrees of freedom.
The n generalized coordinates q i defining any position of the system may be regarded as the
coordinates of a point in an n-dimensional space. We recall:
D

The space of n-tuples of (real) generalized coordinates is called the configuration


space.

Lagrangian Dynamics, Variational Principles

355

This differentiable n-manifold1 is denoted by Q.


Examples.
(i) The configuration space of a system S composed of N particles is R 3 N , each particle
having 3 degrees of freedom.
(ii) The configuration space of a system of p free rigid bodies is R 6 p , each body having 6
degrees of freedom a priori.
(iii) The configuration space of a double pendulum in a plane is a 2-torus, Cartesian product
of two circles S 1 .

(iv) A system of N particles in R 3 , with Cartesian coordinates x, y, z , being subject to


constraints defined by m equations f j ( x, y, z ) = 0 has n = 3 N m degrees of freedom.
In particular, what is the configuration space of a system of two particles linked together by a
rod of negligible mass and moving in R 3 ?
A priori, the two particles have 3 N = 6 degrees of freedom. But an equation of constraint
expressing that the distance between two particles is constant reduces to 5 the number of
degrees of freedom.
Generalized coordinates are the three coordinates of the center of mass G and the two angles
giving the direction of the rod.
The configuration space is R 3 S 2 with generalized coordinates q1 = xG , q 2 = yG , q 3 = z G ,
q 4 = and q 5 = .

Remark. We must note that in any neighborhood of every point of Q, there is a system of
generalized coordinates (q1 ,..., q n ) which is not unique. Other systems of generalized

coordinates exist and can be defined by functions (of class C q , q 1 ):


q ' j = f j (t , q 1 ,..., q n )

j = 1,..., n

such that
D( f 1 ,..., f n )
D(q 1 ,...q n )

0.

Generalized trajectory
D

As a mechanical system S moves, the representative point q = (q 1 ,..., q n ) of this


system describes a path in the configuration space which is called the generalized
trajectory of the system.

For every motion of S, the representative point q must verify the Lagranges equations.
There are many possible generalized trajectories through each point q Q because various
corresponding velocities q 1 ,..., q n (called the generalized velocities) exist a priori.
1

See for instance our book Differential Geometry with Applications to Mechanics and Physics.

356

Chapter 5

Given (q 10 ,..., q 0n ) at the time t 0 , there is one generalized trajectory through q(t 0 ) .
Now, let us express the kinetic energy of S.
The Cartesian coordinates of the respective N particles of S are denoted as follows:
( x1 , y1 , z1 ) = ( x1 , x 2 , x 3 )
.
.
( x N , y N , z N ) = ( x 3 N 2 , x 3 N 1 , x 3 N ) .

The kinetic energy of the system is written:


3N

T = 1 mr (x r ) 2
2

r =1

where the common coefficient m1 = m2 = m3 represents the mass of the first particle and so
on.
Since the various x r are functions of various q j and t, we have:
T=
=

1
2
1
2

x r x r j 2
+
mr (
q )
t q j

3N

r =1

3N

mr (

r =1

x r 2 3 N
x r x r j 1
) + mr
q +
t
2
t q j
r =1

3N

r =1

mr

x r x r
q q
j

q j q k .

The kinetic energy is composed of three terms, namely:


x r 2
) = a (t , q )
T0 = mr (
2 r =1
t
which is independent of generalized coordinates;
1 3N

x r x r j
q = a j (t , q ) q j
j
t q
r =1
which is linear in the generalized velocities;
3N

T1 = mr

T2 =

1 3N

mr

x r x r
k

q j q k = a jk (t , q ) q j q k

q q
which is quadratic in the generalized velocities.
r =1

In the case of scleronomic systems, we have:


x r = x r (q ) ,
x r
= 0 , and the kinetic energy is the following quadratic form in the generalized
t
velocities:

that is

T = a jk (q ) q j q k .
2

(5-23)

Lagrangian Dynamics, Variational Principles

357

This form is assumed positive-definite and the metric element of the configuration space Q is
defined by

ds 2 = a jk (q) dq j dq k = 2T dt 2 .
Example. The configuration space of a system of two mass points linked together by a rigid
weightless rod and moving in a plane is R 2 S 1 .

The equation of constraint expressing that the distance between the two points of respective
masses m1 , m2 and of respective coordinates ( x1 , y1 ), ( x 2 , y 2 ) is the following positive
constant
( x 2 x1 ) 2 + ( y 2 y1 ) 2 = k .

The metric is determined from


1

T = m1 (( x1 ) 2 + ( y1 ) 2 ) + m2 (( x 2 ) 2 + ( y 2 ) 2 ) .

To conclude, we are going to consider the motion of the representative point in the
configuration space of the system S.
Given the previously introduced metric, we state:
PR5

The motion of a mechanical system S is defined by the motion of the representative


point of unit mass under the generalized forces.

Proof. We recall that the representative point must verify Lagranges equations:
d T T

= Qi
dt q i q i

i = 1,..., n

where each Qi is the (covariant) component of the generalized force corresponding to q i ,


called the generalized force associated with q i and such that

W = Qi q i .
The first member of Lagranges equations is written:
1
1
d
(aij q j ) i a jk q j q k = aij q j + k aij q j q k i a jk q j q k
dt
2
2

1
= aij q j + ( j aik + k aij i a jk ) q j q k = aij q j + ijk q j q k
2
j
= aij (q + rkj q r q k ).
We recall that the velocity components of the representative point (called the generalized
velocities) are expressed as follows:
v i = q i

and

358

Chapter 5
vi = aij q j =

q i

The components of the acceleration are

j = q j + rkj q r q k .
Therefore, the Lagranges equations of motion of the representative point are
aij j = Qi

i = Qi

(5-24)

where the various i are the covariant components of the acceleration of the representative
point with unit mass.

1.5

ADJOINT LAGRANGIAN AND FIRST INTEGRALS

Adjoint Lagrangian and generalized momentum


D

The configuration spacetime is the Cartesian product R Q of the configuration


space Q and the time axis R.

From this definition, we consider the Lagrangian


L : R Q R n R : (t , q, q )  L(t , q, q )

which is a function of time, of n generalized coordinates q i and of n generalized velocities


q i such that:
L(t , q, q ) = T (t , q, q ) V (t , q ) .

We multiply each of Lagranges equations by the corresponding generalized velocity and we


sum over i, that is:
q i (

d L L

)=0
dt q i qi

i = 1,..., n

also written:
L
L
d i L
(q
) (qi i + q i i ) = 0 ,
i
dt
q
q
q

that is finally:
d i L
dL L
(q
)( ) = 0.
i
dt
dt t
q

We can express this result in a vivid form by defining the following:

(5-25)

Lagrangian Dynamics, Variational Principles


D

D

359

The generalized momentum or momentum canonically conjugate to the coordinate


q j is
L

(5-26)
pj =
q j
The adjoint Lagrangian is the function
L : R Q R n R : (t , q, q )  L (t , q, q )

such that
L = pi q i L .

(5-27)

In the same manner, it is


L =

T i
L i

q L .
q

L
=
q i
q i

Then, Eq. (5-25) is the so-called complementary equation:


d L
L +
= 0.
t
dt

(5-28)

Jacobis integral

We recall that the first integral terminology designates, in mechanics, a function of


time, positions and velocities which is constant during the motion. We adopt this usual
terminology but we must emphasize that H. Poincar used the more suitable term invariant to
designate this notion.
PR6

Every scleronomic and conservative system has the adjoint Lagrangian as a first
integral called the Jacobis integral.
This first integral is the mechanical energy.

Proof. First, if the Lagrangian does not explicitly depend on time; that is:
L
= 0,
t

then the total derivative


d
L =0
dt

implies that L is a first integral.


Second, by considering
L = pi q i L =

T i
q L ,
q i

then Eulers theorem on homogeneous functions applied to the homogeneous quadratic


function T of generalized velocities (scleronomic case!) is written as follows:

360

Chapter 5
q i

T
= 2T
q i

and thus

L = 2T L = T + V .
Cyclic coordinates
D

A generalized coordinate q k is cyclic (or ignorable or hidden) if it does not explicitly


occur in the Lagrangian:
L
q k

PR7

= 0.

The generalized momentum conjugate to any cyclic coordinate is a first integral of the
equations of motion.

Proof. Given a cyclic coordinate q k , we have:


L
q

=0

d L
=0
dt q k

and thus the first integral is obviously:


pk =

L
=c
q k

(constant).

2. VARIATIONAL CALCULUS AND PRINCIPLES


Extremum principles originated in the following idea: Nature always acts in the
simplest way. They develop many disciplines of physics as geometric optics, mechanics, field
theories, general relativity and so on. For instance, Einstein (1916) wrote1: Recently H.A.
Lorentz and D. Hilbert have succeeded in giving general relativity an especially transparent
form in deriving its equations from a single variation principle. This will be done also in the
following treatment
In all likelihood, the pioneer was Hero of Alexandria (second century BC) which stated the
law of reflection of light, namely: A light ray follows the shortest possible path from one
point to another.
Towards the mid-17th century, this principle of short path was reconsidered by Fermat.
Fermats principle of least time, taking into account the laws of reflection and refraction,
consists in postulating that a light ray follows a path, from one point to another, which
requires the least time.
The Calculus of variations studies extremum principles and is due to mathematicians as
Newton, Leibniz and Jacques Bernoulli particularly.
1

In Hamiltonsches Prinzip und allgemeine Relativittstheorie, Preuss. Akad. Wiss. Berlin, Sitzber., 11111116.

Lagrangian Dynamics, Variational Principles

361

In mechanics, the first development of a variational was introduced by Moreau de Maupertuis


which vaguely formulated the principle of least action in 1747. This principle was based on
no clear notion of action which must be minimized in accordance with theological concepts.
About 1760, Lagrange mathematically expounded the principle of least action and in 1834
Hamilton formulated the general variational principle from which follows mechanics. We will
see that in detail.
First we recall the basic mathematical notion of variational calculus.

2.1

EULERS EQUATIONS

2.1.1

A Variational Problem and Variations

Let1

f : R R U R : ( x, y, y )  f ( x, y, y )

be a function of class C 2 on a definition set { ( x, y ) R 2 } and for every finite value of y (x) .
We consider any curve c defined by the function
y : [ x1 , x 2 ] R R : t  y ( x)

such that
(i)

x [ x1 , x 2 ] : ( x, y ( x)) R R ;

(ii)

y ( x1 ) = y1 , y ( x 2 ) = y 2 , given two reals y1 and y 2 ;

(iii)

Given the rectangle P = { ( , x) R 2 : , x1 x x2 }, there exists a family


of curves such that the functions y ( x) (possibly piecewise of class C 1 on [ x1 , x 2 ] .

Let { c } be such a family of curves.

Fig. 80
We mention that y (x ) denotes the derivative of
will choose the time t as the independent variable.
1

y (x) with respect to an independent variable x. Later we

362

Chapter 5
A functional on { c } is a mapping
{ c } R : y 

x2

x1

f ( x, y ( x), y ( x)) dx

where the integral is called the integral of f along c .


We denote this integral by I (c ) or I c :
I c =

f ( x, y ( x), y ( x)) dx .

(5-29)

Example. The length of a curve is a functional defined by the mapping


{ c } R : y 

x2

x1

1 + y 2 dx .

Remark. As for functions, the image of any curve c under the functional, that is I (c ) , is
sometimes said to be the functional.
The problem consists in finding y which determines c such that I c is an (absolute)
extremum of { I c }. In other words, the problem is to determine y making I c extremum
and fulfilling the limit conditions:

y ( x1 ) = y1 ,

y ( x 2 ) = y 2 .

We mention that we can only search for minima since a maximum for any function is a
maximum for the opposite function.
D

A curve c defined by a function y c makes I c a relative minimum for { I c } if, for


every curve c , there exits R+ such that x [ x1 , x2 ] :

y c ( x) y ( x) <

I c I c

(5-30)

where I c is the integral of f along the curve c making the integral I minimum.
We specify that any curve making an (absolute) extremum automatically makes a relative
extremum; so, afterwards the extrema will be understood as relative ones.
Now, we set

y ( x) = y c ( x) + ( x)
also denoted y ( , x) and where (x) is a function of class C 1 such that ( x1 ) = ( x 2 ) = 0 .
The parameter (real variable) is sufficiently small in order that the condition (5-30) be
fulfilled.
The family of functions y is arbitrary and there are many others.
Notation. We simply denote y c ( x) by y (x) .

363

Lagrangian Dynamics, Variational Principles

The integral of f along c:


Ic =

x2

f ( x, y ( x), y ' ( x)) dx

x1

corresponds to the case where = 0 .


D

The complete variation of y is

y ( x) = y ( x) y ( x)
= ( x).

(5-31)

The problem is thus to look for y ( x) such that

y ( x ) <
with

y ( x1 ) = y ( x 2 ) = 0 .
The integral I c is function of the chosen and we denote simply:
I ( ) =

x2

I ( ) =

x2

x1

f ( x, y ( x), y ( x)) dx

(5-32)

also denoted
x1

f ( x, y ( , x), y ( , x)) dx

and obviously

I (0) = I c .
D

The complete variation of I is the increase


I ( ) I c =

x2

x1

f ( x, y + y, y + y ) dx

x2

x1

f ( x, y, y ) dx .

For sufficiently small values of we have:


I ( ) I (0) .

The conditions of minimum of I ( ) , at = 0 , are


I (0) = 0

I (0) 0 .

Let us calculate I ' (0) .


Since f is of class C 2 , we have:

I ( ) =

d
d

x2

x1

which implies

x2
1

f ( x, y + , y + ) dx

f ( x, y + , y + )
f ( x, y + , y + )
) dx
+
( y + )
( y + )

364

Chapter 5

I (0) =

x2

x1

f
f
( x, y, y ) +
( x, y, y )) dx = 0 .
y
y

If y is of class C 2 , we have:
x2

x2
d f
f x2
f
dx = [
] x1 (
) dx
x1
dx y
y
y
x2
d f
dx
=
x1
dx y

and the condition of extremum I (0) = 0 becomes:


x2

I (0) = (
x1

f
d f

) dx = 0
y dx y

for every function of class C 1 vanishing at x1 and x2 .


D

The (first) variation of the integral I is I (0) , that is


x2

I = y (
x1

f
d f

) dx ,
y dx y

(5-33)

where we recall y = (x) .


We can say:
PR8

A necessary condition for the integral I ( ) to be an extremum for y = y (x) is

I =0.

(5-34)

Remark. We can interpret y = as an infinitesimal variation of the function defining a


curve of { c } and making I extremum ( I = 0) ; that is, leading to a higher order variation
of integral I.

2.1.2

Eulers Equations

Let us first establish the famous Euler equation.


PR9


The condition I = 0 of cancellation of the (first) variation of I is a necessary and


sufficient condition of existence of Eulers equation

d f
f

= 0.
y dx y

(5-35)

Proof. The only difficulty is to prove that Eulers equation ensues from I = 0 ; but it is
immediate from the following lemma:

Lagrangian Dynamics, Variational Principles

365

If a continuous function

g : [ x1 , x 2 ] R : x  g ( x)
fulfills the following condition
x2

( x) g ( x) dx = 0

for every (x) of class C 1 on [ x1 , x 2 ] vanishing at x1 and x2 , then g = 0 on [ x1 , x 2 ] .


Indeed, let us assume that g is strictly positive for every x ]x1 , x 2 [.
Since g is continuous on ]x1 , x 2 [ , it will be strictly positive on a neighborhood of x :
] x a, x + a [ ] x1 , x2 [ .

In this neighborhood, there is thus a positive real c such that:


g ( x) > c .

Given any function of class C 1 , positive on ] x a, x + a [ and vanishing elsewhere, but


more particularly equal to unity on ] x a 2 , x + a 2 [ , we have:
x2

x + a 2

( x) g ( x) dx =

x a 2

g ( x) dx c a > 0

which is at variance with the assumption.


Thus it is absurd to assume that g is positive for every x ]x1 , x 2 [ .
A similar conclusion would be obtained by choosing a strictly negative function.
So g (x) vanishes on ]x1 , x 2 [ and the lemma is proved.

By applying this lemma to the necessary condition (5-34) of extremum for I, for which the
lemma hypotheses are fulfilled, we obtain the famous Euler equation of variational calculus.
By denoting

f f d f
=

y y dx y

the so-called variational derivative of f, then Eulers equation is written:

f
= 0.
y
We emphasize that every function y of class C 1 , introduced as before, making I extremum,
must necessarily verify the Euler equation.
We say:
D

A curve x  y (x) solution of Eulers equation is called an extremal.

An extremal does not necessarily produce an extremum since it fulfills only a necessary (and
not sufficient) condition. For instance, saddle points are possible.

366

Chapter 5

First integrals

Eulers equation is a second-order differential equation, namely:


f
2 f
2 f
2 f

= 0.
y
+
y
+
2
y y
x y y
y

(5-36)

It defines functions depending on two arbitrary constants:


y = y ( x; c1 , c 2 ).
Any problem of the variational calculus is obviously different from a (local) Cauchys
problem because it consists in obtaining an extremal through two points (two arbitrary
constants being determined by limits).
Even for simple functions, the problem of existence and unicity of solutions is not always
obvious! It is possible that there are one or several extremals through two limit points or
none!
But the integration of the Euler equations is simplified if there exists first integrals. So,
(i)

If f does not explicitly depend on y, then

f
d f
= 0 and thus
is a first integral of
dx y
y

Eulers equation.
(ii)

f
is a first integral since
y
f
f
f
f
d f
d
y +
y y
( f y ) =
y
= 0.
y
y
y
y
dx
dx y

If f does not explicitly depend on x, then f y

System of Eulers equations

The variational method is immediately generalized to an integral of type


x2

f ( x, y i ( x), y i ( x)) dx

where f is a C 2 function of 2n + 1 variables x, y i , y i .


The various y i belong to the space of C 2 real functions (or piecewise C 1 ) defined on
[ x1 , x 2 ] R .
By analogy with previous developments, we let:
yi ( x) = y i ( x) + i ( x)

i = 1,..., n ,

y i = yi ( x) y i ( x)
where y i ( x) = y 0i ( x) .
It is easily proved that the (first) variation of I ( ) is (sum over i):
x2

x1

I = y i (

d f
) dx .
dx y i

(5-37)

Lagrangian Dynamics, Variational Principles

367

The following is immediately proved.


A necessary and sufficient condition for a curve of equations y i = y i (x) fulfills the
necessary condition of extremum I = 0 is provided by the following system of Eulers
equations:

f
f
d f
) i
= 0.
i
dx y i
y
y

i = 1,..., n .

(5-38)

We note this is a system of n second-order differential equations with 2n conditions


y i ( x1 ) = a i

y i ( x2 ) = b i

where a i and b i are 2n given reals.


Example of the brachistochrone problem (Jacques Bernoulli).

A wire in a vertical plane links up two points a and b. A particle of mass m, at rest at a, slides
without friction down the wire to b under gravity. Find the equation of the curve for which the
particle goes from a to b in the least time.
Answer. The well-known relation
dT = dW

is expressed as
1

d ( mv 2 ) = mg . ds
2

and the initial conditions imply:


1
2

mv 2 = mg k . ds = mg z ,
ab

k being downwards directed.


Thus the velocity of the particle is
ds
= 2g z
dt

which implies
1 + (dz dx) 2

dt =

2g z

dx .

The problem consists in minimizing


1

1 + z2

xb

2g

dx .

Since the integrand does not explicitly depend on x, a first integral is


1 + z 2

z 2
z 1 + z2

=c

368

Chapter 5

or
z 1 + z 2 =

= k.

Thus we have:
dx =

z
dz .
kz

By letting
z = k sin 2 =

k
2

(1 cos 2 ) ,

we have:
x = 2k sin 2 d =

k
2

(2 sin 2 ) + K .

Since the curve passes through the origin, we must have K = 0 , so that the required equations
are (setting R = k 2 ):
x = R (2 sin 2 )

z = R (1 cos 2 ) ,

the constant R being determined from the limit condition at b.


The extremal is a cycloid, that is the path of a fixed point on a circle of radius R which rolls
along the x-axis.

2.2

HAMILTONS VARIATIONAL PRINCIPLE

2.2.1 Hamiltons Postulate

We know that dAlembert-Lagrange principle and (infinitesimal and instantaneous)


virtual displacements show a differential character and let obtain the famous Lagrange
equations (1788). However, these equations of motion can be deduced from a variational
principle due to Hamilton (1834).
If we substitute L(t , q i , q i ) for the function f ( x, y i , y i ) introduced in the variational
calculus, we conclude that the Euler equations of the calculus of variations become the
Lagrange equations obtained from the nonvariational principle of dAlembert-Lagrange. This
remarkable observation is at the root of Hamiltons principle in dynamics.
Let us introduce an essential definition in order to present this principle.
We consider two fixed points q1 and q 2 of the configuration space Q and the following set of
curves of class C 2
E = { c : [t1 , t 2 ] R Q ; c(t1 ) = q1 , c(t 2 ) = q 2 }.

The function S : E R is defined by the action integral


t2

S (c) = L(t , c(t ), c(t )) dt .


t1

(5-39a)

369

Lagrangian Dynamics, Variational Principles

The action integral is also denoted by


t2

S = L(t , q i , q i ) dt .
t1

(5-39b)

Now, we can state the Hamiltons variational principle:


PR10 Among all the possible paths along which a material system can move from an instant
t1 to an instant t 2 , the actual path is such that the (first) variation of the action integral
vanishes:


S =

t2

L(t , q i , q i ) dt = 0 .

(5-40)

In other words, a motion in the configuration space Q is a mapping


R Q : t  (q 1 (t ),..., q n (t ))

making extremum the action integral.


The general nature of this variational principle (viewed as a postulate) widens the field of
theory compared with Newtons laws of motion.
Hamiltons variational principle is also denoted

S =

2.2.2

t2

( pi q i L ) dt = 0 .

(5-41)

Hamiltons Principle and Motion Equations

We know that dynamics can be considered in two opposite manners.


In the first, a variational principle is assumed and Newtons or Lagranges equations are
derived as theorems.
In the second, the Lagrange (or Newton) equations are taken as axioms and variational
principles are deduced as theorems.
We are going to prove implications between Hamiltons principle and motion equations.
PR11 The extremals relating to Hamiltonians variational principle verify the corresponding
Newton equations.
Proof. Let us consider a system of N particles of position vectors rh , of corresponding
masses mh and without constraints. The extremals make the following integral extremum:
t2

t2

t1

t1

S = L dt = (
t2

= (
t1

3N

mh rh2 V (r1 ,..., rN )) dt

h =1

mr ( x r ) 2 V ( x1 ,..., x 3 N )) dt ,

r =1

written with the notations of Section1.4.

370

Chapter 5

The extremals are solutions of 3N Eulers equations:


d L L
d T V
= 0.
+
i =
i
dt x
dt x i x i
x

These equations, grouped together 3 by 3, are really the equations of Newton for each of
particles:
M 1r1 + r1V = 0 , . . . , M N rN + rN V = 0

where
m1 = m2 = m3 = M 1 , . . . , m3 N 2 = m3 N 1 = m3 N = M N .
PR12 The motion of the representative point of a system takes place in the configuration
space iff the generalized velocities of this point satisfy the Lagrange equations.
Proof. By analogy with (5-32), we have:
t2

S ( ) = L(t , qi (t ), qi (t )) dt
t1

and also
t2

S (0) = L(t , q i (t ), q i (t )) dt .
t1

Hence we have:
t2
L dq i
L dqi
d
) dt
S ( ) = ( i + i
t1
d
q d q d
t2

= (
t1

t2

t1

i = 1,..., n

L dqi d L dqi
d L dqi
(
)
(
)
) dt
+

dt qi d
qi d dt qi d

L
d L dqi
( i
)
dt
q dt qi d

since there is no variation at the limits.


Thus we have the (first) variation:
t2

t1

S ( ) =0 = (

d L
) i (t ) dt = 0
dt q i

also written [see Eq. (5-37)]:


t2

t1

S = (

d L
) q i dt = 0 .
i
dt q

From a previous lemma of the calculus of variations and since the various q i are arbitrarily
independent, then the last integral is zero iff i { 1,..., n }:
d L L

=0.
dt q i q i

Lagrangian Dynamics, Variational Principles

371

Remark. The reader will also refer to Exercise 8 where Hamiltons principle is deduced from
Lagranges equations and the complementary equation.

We must also mention the following


PR13 Given a gauge transformation of a Lagrangian, that is
L (t , q, q ) = L(t , q, q ) + df (t , q )

where f is a function of class C 2 of variables t and q i , then the extremals of EulerLagrange equations correspond for L and L .
In other words, any extremal defined by q (t ) is such that

L L d L
) i
=0
dt q i
q
qi

L
L d L
) i
= 0.
i
dt q i
q
q

We emphasize that f does not depend on q i .


Proof. The various variational derivatives are

L
L
df
d df

L
=
+
=
+( i
(
)
)( )
i
i
i
dt
dt q i dt
q
q
q
q
L d f
f
f
=
+ ( i i ( + k q k ))
i
q dt q q t q
L
=

qi

So we conclude as follows:

L
=0
qi

2.3

L
=0
qi

i = 1,..., n.

JACOBIS FORM OF THE PRINCIPLE OF LEAST ACTION

The following presentation of the principle of least action shows geodesics as


extremals of arc length and permits bringing closer mechanics and geometrical optics.
We consider a scleronomic system S of which the potential is V (q i ) .
The generalized forces associated with the n generalized coordinates q i are
Qi =

V
q i

We also recall the expression of the kinetic energy


1

T = aij q i q j .
2

372

Chapter 5

PR14 The generalized trajectories of the representative point of the motion of S in the
configuration space and admitting the first integral1 E = T + V are the geodesics
corresponding to the metric
d 2 = 2T aij dq i dq j
= 2( E V ) aij dq i dq j

(5-42)

also written:
d 2 = (2U + h) aij dq i dq j .

(5-43)

The representative point follows the geodesics according to the time law
d
= 2U + h .
dt

(5-44)

The generalized trajectories are the extremals corresponding to


t2

2T dt = 0 .
t1

(5-44)

Proof. The Lagranges equations are


V
d
1 a jk j k
q q = i
(aij q j )
i
dt
2 q
q

(5-46)

We recall that the kinetic energy defines a metric on Q:


ds 2 = 2T dt 2 = aij dq i dq j .
We choose for the configuration space a new metric:
d 2 = F ds 2 = F aij dq i dq j
where, a priori, F is a strictly positive function of q i so that the generalized trajectories of the
representative point are geodesics and where the parameter is a function of t.
On the one hand, if we multiply the members of the motion equations (5-46) by

dt
, these
d

equations become:
d
dq j d
1
dq j dq k d
dt
=
iV .
(aij
) i a jk
d
d dt
2
d d dt
d

(5-47)

On the other hand, for this metric, the equations of geodesics of the configuration space are:
j
k
d 2qi
i dq dq
+

=0
jk
d d
d 2

where the Christoffel symbols correspond to the line element d 2 = ( F aij ) dq i dq j .


From this element, the equations of geodesics are written with covariant components as
follows:
1

Or in an equivalent manner : h = 2(T U ), U = V being the force function.

Lagrangian Dynamics, Variational Principles


j
k
d 2q h
h dq dq
+ jk
)
F aih (
d d
d 2
d 2q h
dq j dq k
= F aih
+ ijk
d d
d 2
h
dq
dq h d
d
=
( F a ih
)
( F aih ) + ijk
d
d d
d
dq k
dq j
d
=
( F a ij
) + ( ijk k ( F aij ))
d
d
d

373

dq j dq k
d d
dq j
= 0.
d

But, since
1

ijk = [ j ( F aik ) + k ( F aij ) i ( F a jk )]


2

the previous equations become


1
d
dq j
dq j dq k
( F aij
) + [ j ( F aik ) + k ( F aij ) i ( F a jk ) j ( F a ik ) k ( F aij )]
=0,
d
d
2
d d

that is:
d
dq j
F
dq j dq k 1
dq j dq k
( F aij
) i a jk
a jk i F
=0
d
d
2
d d 2
d d

or
d
dq j
F
dq j dq k 1 i F
( F aij
) i a jk
=

d
d
2
d d
2 F

(5-48)

By comparing Eqs. (5-47) of the motion of the representative point with Eqs. (5-48) of
geodesics, we adopt the following
F=

d
dt

and thus for this choice, we have:


i F = 2 i V .
By integrating, we obtain
F = 2 V + C

where C is a constant.
From
d 2 = F aij dq i dq j =

and thus

d = 2T dt

we deduce the following expression


F = 2T .

d
2T dt 2
dt

374

Chapter 5

Therefore, the constant is


2T + 2V = 2 E

which is often written with the function force U as follows:


2T = 2U + h .

We have so obtained the extremals as geodesics corresponding to the metric defined by


d 2 = (2U + h) aij dq i dq j .
The corresponding variational principle is written:

t2

2T dt = 0 .

We specify that the representative point follows every corresponding generalized trajectory
according to the time law defined by
d
= 2U + h
dt

that is
dt =

ds
2T

ds
2U + h

3. EULER-NOETHER THEOREM
Let us introduce a general theorem which plays a fundamental role in Lagrangian
mechanics and that Euler used already under another form (obviously). This so-called EulerNoether theorem lets us associate a first integral of Lagranges equations to any oneparameter group of diffeomorphisms (on the configuration space) which conserves the
Lagrangian.
Given U and U opens of the configuration space Q, we recall:
D

A mapping f : U U is a diffeomorphism of class C r if f is a bijection of class C r


such that f

3.1

is of class C r .

ONE-PARAMETER GROUP OF DIFFEOMORPHISMS


We consider a material system with n degrees of freedom.

Let J be an interval of R,
L : U R n : (q, q )  L(q, q ) be a (differentiable) Lagrangian.
D

A tangent vector field X on Q is a mapping which assigns a pair (q, q ) to each point
q Q :
X (q ) = (q, q ) .

Lagrangian Dynamics, Variational Principles


D

375

An integral curve of field X on Q is a curve


c : J Q : t  c(t )

such that
d c(t )
= X (c(t )) .
dt

Given any field X and any point q Q , we consider:


-

a neighborhood U of this point,


an interval I = { t R : t < , R+ },
a differentiable mapping:

: I U Q : (t , q)  (t , q) = t (q)
such that

0 (q ) = q .
We introduce the following definition:
D

A local transformation of Q generated by a tangent vector field X is a diffeomorphism


between neighborhoods of Q:

t : U Q : q  t (q )
verifying the following differential equation
d
t (q) = X ( t (q))
dt

(5-49)

and

0 (q ) = q .
Now, we can introduce the notion of one-parameter group of diffeomorphisms and prove the
following
PR15 Every tangent vector field on Q generates a one-parameter group of diffeomorphisms
on Q.
Proof. Let I = { t : t < , > 0 } and be a neighborhood of {0} Q . We consider a local
transformation
U Q : q  t + s (q )
with reals t , s, t + s I .
This local transformation satisfies, such as t (q) , the differential equation associated with the
field X.
The following integral curves associated with X
t  t + s (q)

t  t ( s (q))

having for t = 0 the same value s (q) [because 0 ( s (q)) = s (q ) ], are such that:

376

Chapter 5

t + s (q) = ( t  s )(q) ,
for every t , s, s + t I (see uniqueness theorem).
Therefore, the local transformation of Q are such that:

t + s = t  s .
In particular, we deduce:

s = s1
because
( s  s )(q) = s + s (q ) = 0 (q) .
In conclusion, the local transformations of Q have a group structure with composite law, 0
being the identity element. This group is called the one-parameter group of diffeomorphisms.

3.2

EULER-NOETHER THEOREM

Before stating the present Noethers theorem, we say:


D

A diffeomorphism s is admissible for Q with a Lagrangian L(q, q ) if the Lagrangian


is invariant under the local transformation:

s : U Q : q  s (q ) .
Example. Given

Q 2 = { ( x , y ) : x R, y R }
and
L=

m
2

( x 2 + y 2 ) V ( y ) ,

then a translation of x, namely s I R :

s : Q2 Q2 : ( x , y )  ( x + s , y )
is obviously admissible.
Now, let us prove the theorem:
PR16 If a one-parameter group of diffeomorphisms is admissible for a configuration space Q
[that is: L(q, q ) invariant under various s ], then there is, at least, a first integral of
Lagranges equations.
Proof. Let c : R Q : t  c(t ) be an integral curve of Lagranges equations.
Since s is admissible for Q (with L), then every transformed curve s  c is also solution of
Lagranges equations. Explicitly, by considering
: R R Q : ( s, t )  ( s, t ) = s (c(t )) ,
we know that

377

Lagrangian Dynamics, Variational Principles

s  c : R Q : t  s (c(t ))
verifies the equations of Lagrange too. We have thus:
d L
(( s, t ), ( s, t )) =
(( s, t ), ( s, t ))
i
dt

q
q
L

i = 1,..., n .

(5-50)

Fig. 81
Since s is admissible, we have [letting q i = i ( s, t )] :
d
L i L  i

L(( s, t ),( s, t )) = i
+
=0.
ds
q s q i s

By taking Eq. (5-50) into account, the last sum of 2n terms becomes:
d L i L d i
d L i
0
(
).
=
=
+
dt q i s q i dt s
dt q i s

So, at point c(t ) , the first integral is denoted by

L d (q)
q ds s

s =0

by denoting the derivative with respect to s by q , this first integral is simply

L
q .
q

(5-51)

Remark 1. For a nonautonomous material system with a Lagrangian L(t , q, q ) , we consider


the configuration spacetime R Q .
As in Exercise 8, we view that the 2(n + 1) dimensional space of points and directions
associated with R Q and the elements of which are
( z j (u ), z j (u ))

such that
- for j = 0 :
- for j 0 :

dt
= t ,
du
dq i
z j (u ) = q i (u ) , z j (u ) =

du

z 0 (u ) = t (u ) , z 0 (u ) =

(arbitrary parameter u)

378

Chapter 5

We know that the action integral is


t2

u2

t1

u1

S = L(t , q, q ) dt =
=

u2

u1

L (t (u ), q(u ),

dq
du

dt dt
) du
du du

L( z j (u ), z j (u )) du

where
dt
du

L=L

and so this case is related to the previous study.


Remark 2. We mention that if L does not explicitly depend on t (that is, the one-parameter
group of time translations s (t , q) is admissible), then there is a first integral that is the
mechanical energy.
We conclude this section with two classical examples illustrating the previous essential
theorem [see, e.g. Arnold (1978)].
Example 1. If the Lagrangian of a constrained material system is invariant under translations
along an axis, defined by e1 for instance, then the first component of the linear momentum is
a first integral.
Answer. The Lagrangian of the system is
L=

mh rh2 V (r1 ,..., rN ) .

h =1

The invariance under any translation, for instance:

s : rh  s (rh ) = rh + s e1
implies
d
s (rh )
ds

s =0

= e1

and the Euler-Noether theorem leads to the following first integral


N

h =1

mh rh . e1 = mh x h
h =1

which is the first component of the linear momentum of the system.


Example 2. If the Lagrangian of a material system is invariant under rotations about an axis,
defined by e1 for instance, then the angular momentum about this axis (or projection of L
onto e1 ) is a first integral.
Answer. The invariance under the above-mentioned rotation

s : rh  s (rh ) = rh
defined by

Lagrangian Dynamics, Variational Principles

379

0
0 xh
xh 1


y h = 0 cos s sin s y h
z 0 sin s cos s z
h
h
implies
d
rh
ds

s =0

= z h = rh e1 .
y
h

The Euler-Noether theorem leads to the well-known first integral


N

h =1

mh rh . (rh e1 ) = (mh rh rh ) . e1 = L . e1 = L x .


h =1

4. EXERCISES
First courses of theoretical mechanics deal with simple Lagrange equations; it is the
reason why we will essentially consider Lagrange equations with multipliers.
Exercise 1.

Two particles of same mass m move along a circle of center o and radius R. They repel
each other under the action of a force inversely proportional to the cube of their distance.
Friction being negligible, prove that the middle c of the arc joining positions a and b of the
respective particles is moving in a circle at constant velocity.
Answer. The middle c of arc ab has one degree of freedom. The chosen generalized
coordinate is the angle locating the position vector oc.
If denotes the angle (oa , oc ) , the values of at a and b are respectively:

a = ,

b = + .

The kinetic energy of the system of two particles is


T=

m
2

R 2a2 +

m
2

R 2b2 = mR 2 ( 2 +  2 ).

Let us set
r = ab = r 1r ,

r= r .

The force exerted on a is


fa =

k
k
1 = 3 1r
3 ba
r
r

( k R+ ).

The work done by forces during differential displacements of a and b is

380

Chapter 5

dW = f a . d oa + f b . d ob =
=

k
1r . d (oa ob)
r3

k
k
1 . d (r 1r ) = d ( 2 ) = dV .
3 r
2r
r

The potential is thus expressed as

V=

k
k
=

2
2
2r
8 R sin 2

Since
L = mR 2 ( 2 +  2 )

k
,
8 R sin 2
2

the Lagranges equation:


d L L
=0

dt 

is reduced to
d L
= 0.
dt 

Thus the constant


L
= 2mR 2


proves that the circular motion of c is uniform.

Exercise 2.

A quarter of a homogeneous circular rod S of radius R and mass M is thrown in a


vertical plane. Find Lagranges equations.
Answer.

Let { o; e1 , e 2 } be a frame of reference. The moving rod has three degrees of

freedom, the chosen generalized coordinates are the two coordinates q 1 = x, q 2 = y of G and
the inclination angle q 3 = of a straight line fixed in S.
The kinetic energy of S is
1
1
1
T = aij q i q j = M ( x 2 + y 2 ) + I 2
2

where I is the moment of inertia about the instantaneous rotation axis; namely:

I = MR 2 .
We have thus
a11 = M ,

a 22 = M ,

a33 = MR 2 ,

( aij = 0 i j ).

381

Lagrangian Dynamics, Variational Principles

The force exerted on the rod, that is the weight f = Mg e 2 , is derivable from a potential
expressed as
V = Mg y .
The Lagrangian is thus
L=

1
aij q i q j Mg q 2 .
2

From
L
= a11 q 1 ,
x

L
= 0,
x

we deduce the following Lagranges equation:


M x = 0 .

(1)

From
L
= a 22 q 2 ,
y

L
= Mg ,
y

we deduce another Lagranges equation:


M y + M g = 0 .

(2)

From
L
= a33 q 3 ,


L
= 0,

we deduce the last Lagranges equation:


1
2

MR 2 = 0 .

(3)

Exercise 3.

A particle of mass m is constrained to move on the inside surface of a smooth


paraboloid of equation z = x 2 + y 2 . By considering one multiplier, find Lagranges equations
and give the physical meaning of the multiplier.
Answer. Given primitive cylindrical coordinates
u1 = r ,

u2 = ,

u3 = z ,

the Lagrangian of the moving particle is


L=

m
2

(r 2 + r 2 2 + z 2 ) mg z .

The constraint equation


r2 z = 0

implies

2r r z = 0 .

382

Chapter 5

By referring to the general constraint equation (5-18), we have:


a11 = 2r ,

a12 = 0 ,

a13 = 1 .

The Lagrange equations with one multiplier are thus


d
dt
d
dt
d
dt

L L

= 2 r ,
r r
L L

= 0,

L L

= .
z z

Thus we have
m(r r 2 ) = 2 r ,
d 2
(r ) = 0 ,
dt
m z = mg .
m

Since in cylindrical coordinates we know that


m d 2
ma = m(r r 2 ) 1r +
(r ) 1 + mz 1z ,
r dt

we have here:
ma = 2 r 1r mg 1z 1z .

So, from
ma = mg + L ,

we deduce the following force of constraint


L = (2r 1z )
and the z-component is such that:

= Lz .
Exercise 4.

Find the law of motion of a particle of Lagrangian


L=

m 2
( x + y 2 )
2

and subject to a constraint of equation


y = x

(or dy x dt = 0 ).

Give the physical interpretation of the introduced multiplier.


Answer. In this problem with one degree of freedom, x and y are primitive coordinates and
we introduce one multiplier 1 .
The equation of constraint y = 0 corresponds to a11 = 0 and a12 = 1 in the general equation
a1i u i = 0 . The Lagrange equations with multiplier are

Lagrangian Dynamics, Variational Principles

383

my = 1 .

mx = 0 ,

The motion of the particle is known by adding the equation of constraint:


y = x .

Indeed, given initial values x0 , y 0 , x 0 , y 0 we have:


x = x 0 t + x0 ,
y = x 0

t2
2

+ x0 t + y 0 .

The Lagrange multiplier

1 = m y = m x 0
is the y-component of the force of constraint since:
r
= L x = 1 a11 = 0 ,
x
r
L . = L y = 1 a12 = 1 .
y
L.

Exercise 5.
A vertical hoop (with its axis always horizontal) of mass M and radius R is rolling
without slipping down an inclined plane on a distance l, denoting the inclination angle.
Express the friction force from Lagranges equations with multiplier.
Answer. The moving hoop has one degree of freedom. We choose two primitive coordinates:
-

the coordinate x which measures the distance covered by the hoop moving down the
plane,
the coordinate which is the angle of rotation about the horizontal axis of the hoop.

The equation of constraint is


dx = R d .

The kinetic energy of the moving hoop is the sum of the kinetic energy about the mass center
G of the hoop and the kinetic energy of rotation about G:
T = M x 2 + MR 2 2 .
1

The Lagrangian is immediately


L=

M
2

( x 2 + R 2 2 ) Mg (l x) sin .

The constraint implies:

x R = 0
and thus
a11 = 1 ,

a12 = R .

384

Chapter 5

Thus Lagranges equations with multiplier are


d
dt
d
dt

L L

= ,
x x
L L

= R ,


that is:
M x Mg sin = 0 ,

(1)

MR + = 0 .

(2)

We add the equation of constraint


x = R ,

(3)

which allows the determination of unknowns , x and .


So, along the x-direction, the motion equation is immediately:
M x M

g
2

sin = 0 .

From
M x = Mg sin + L x = M

g
2

sin ,

where L x is the constraint force of friction, we deduce:


g
L x = = M sin .
2

The equation of motion shows that the hoop is rolling with half the acceleration of the
corresponding frictionless motion.
Exercise 6.

A small ring p of mass m frictionless slides along a circular hoop of radius R which
rotates about a vertical axis at a constant angular velocity  = . This example is the one of
Section 2.2.3.d of Chapter 1.
(i) Find the equation of motion of p by introducing an obvious generalized coordinate.
(ii) Determine the components of the force of constraint exerted on p by considering the
spherical coordinates r , , . Give the physical interpretation of Lagrange multipliers for
these primitive coordinates.
Answer. (i) Given the generalized coordinate which is the angle (k , r ) where r is the
position vector of p, we have immediately:
L=

m
2

R 2 ( 2 + 2 sin 2 ) mg R cos

and thus the motion equation of Lagrange is


R R 2 sin cos g sin = 0.

385

Lagrangian Dynamics, Variational Principles


(ii) Let us introduce the (primitive) spherical coordinates u 1 = r , u 2 = and u 3 = .

This problem has one degree of freedom and there are two equations of constraint.
The first one is
rR=0
( r = 0 ),
that is

a1i u i = 0
with a11 = 1 , a12 = a13 = 0 .

The second equation of constraint is

t = 0

( = 0 ),

that is
a 2i u i = 0

with a 21 = a 22 = 0, a 23 = 1 .
Lagranges equations with multipliers are:
-

First:
2

mr mr ( 2 +  2 sin 2 ) + mg cos = r a r1

(1)

r =1

where the second member, equal to 1 , is the generalized force of constraint associated with
u 1 , namely:

1 = L
-

r
= L . 1r = Lr .
r

Second:
m

2
d 2
(r ) mr 2 sin cos  2 mgr sin = r a r 2
dt
r =1

(2)

= 0.

Third:
2
d 2
2

m (r sin ) = r a r 3
dt
r =1

(3)

where the second member, equal to 2 , is the generalized force of constraint associated with
u 3 , namely:

2 = L

r
= L . r sin 1 = r sin L .

By taking into account of two constraints ( r = R, = t ), we notice that Eq. (2) is the
equation of motion already found in (i).
Given these constraints, Eq. (1) expresses the radial component of the force of constraint on p:
Lr = mR ( 2 + 2 sin 2 ) + mg cos

which is 1 .

386

Chapter 5

In the same conditions, Eq. (3) expresses the longitudinal component of the force of constraint
on p:
L = 2m R cos .
The multiplier 2 is L multiplied by R sin .
Exercise 7.

The extremity a of a homogeneous rigid rod ab is confined to move along a fixed


vertical axis, when the end b is confined to move on the plane through o perpendicular to axis.
The motion of the rod of mass m and length l is assumed frictionless.
(i) Determine Lagranges equations of motion.
(ii) Find first integrals and give their interpretation.
Answer. (i) The extremity a has one degree of freedom, while b has two degrees of freedom
a priori; but the distance between a and b is constant, which introduces a constraint. So, the
rod has two degrees of freedom.

Fig. 82

Let R e = { o; e1 , e 2 , e3 } be a frame of reference such that e3 is vertical.


The position of the rod is determined with respect to R e if the angle of position of the
vertical plane oab is known with respect to the vertical plane { o; e 2 , e 3 } and if the angle of
position of the bar in the plane oab is known with respect to the vertical axis.
The angular velocity vector of the rod (with respect to R e ) is obviously
=  e 3 +  E1

where E1 is the unit vector perpendicular to the plane { o; e 2 , e 3 } of the rod as shown in the
previous figure.
We are going to calculate the kinetic energy, namely:

387

Lagrangian Dynamics, Variational Principles


T =

1
2

m vG

+ I ij(G ) i j
2

where v G is the velocity of the mass center G with respect to R e and the various I ij(G ) are
the components of the inertia tensor of the rod expressed in frame { G; E1 , E 2 , E 3 } fixed in
the moving rod.
First, time derivatives of oG with respect to R e and R e = { E1 , e 2 , e 3 } are connected as
follows:
de
d e
vG =
oG =
oG + ee oG
dt
dt
d e
=
oG +  e 3 oG.
dt
But, from
l
d e
l
oG =  cos e 2  sin e 3
dt
2
2
l
l
(since oG = sin e 2 + cos e 3 ),
2

and from
l

 e 3 oG =  sin E1 ,
2

we deduce:
l
l
l
v G =  sin E1 +  cos e 2  sin e 3
2
2
2

and thus
v G2 =

l 2 2 2
( + sin 2 ) .
4

We mention that the previous result is also obtained from


de
l
oa + E 3 ( e 3 +  E1 )
dt
2
l
l
= l sin e 3  sin E1 +  E 2

v G = v a + Ga =

since
E 2 = cos e 2 + sin e 3 .
Second, since e 3 = sin E 2 + cos E 3 , we have immediately:
ml 2 12 0

1 (G ) i j
1  

I ij = ( sin cos ) 0
ml 2 12
2
2

0
0

ml 2  2  2
( + sin 2 ).
24

0 

0  sin

0  cos

388

Chapter 5

So, the kinetic energy of the moving rod is


l 2 2 2
T = m ( + sin 2 )
6

and the Lagrangian is


l 2 2 2
l
L = m ( + sin 2 ) mg cos .
6

The Lagrange equations are immediately:


l2
ml 2  2
l
d
(m  )
sin cos mg sin = 0 ,
3
3
2
dt
2
d
l
(m  sin 2 ) = 0
dt
3

or explicitly:

  2 sin cos

3 g
2 l

sin = 0 ,

 sin 2 + 2 sin cos = 0.


(ii) The constraint being frictionless, the force mg is derivable from a potential V and the
mechanical energy is a first integral, namely:
T +V = m

l 2 2 2
l
( + sin 2 ) + mg cos .
6

Another first integral is obviously  sin 2 = c , that is 0 sin 2 0 given initial angles 0 and
0 .
Let us express this first integral in function of angular momentum La .
d
La is the moment about a of external forces plus
From Eq. (4-23), we know that
dt
mv G v a .

The external forces are mg and the forces of constraint La and Lb .


We know that M a ( La ) = 0 ; M a (mg) and M a ( Lb ) are orthogonal to axis defined by e3 , so
the projections of these moments on this axis are zero. In addition, the projection of mv G v a
on this axis is obviously zero.
Thus we conclude that
d
e3 La = 0
dt
which implies:
e3 . La = k
(constant).
Let us calculate the following
La = m aG v a + I ( a ) .
First, we have:

Lagrangian Dynamics, Variational Principles


l

389

l
m aG v a = m E 3 (l sin e 3 ) = m  sin 2 E1 .

Second, from the inertia tensor I (G ) and Steiners theorem, we deduce, with respect to
{ a; E1 , E2, E3 }, the following
Ml 2 3
0
0 

I (a) = 0
Ml 2 3 0  sin

0
0  cos
0

and thus
l2
 
La = ml 2 ( sin 2 ) E1 + m  sin E 2
3

and finally
La . e 3 = m

l2  2
sin = k .
3

So, the connection between the first integral 0 sin 2 0 and the projection of La on the
vertical axis is established.
Exercise 8.

By considering the configuration spacetime of a system S, deduce Hamiltons


variational principle from Lagranges equations and the complementary equation; these
equations correspond to the natural trajectory described by a material system S from one
point (t1, q 1 ) to another (t 2 , q 2 ).

Answer. We know that the motion of a material system of n degrees of freedom is described
by successive n-tuples q (t ) = (q1 (t ),..., q n (t )) of Q.
We emphasize that the variables q i and t are treated on an equal footing and there are so
expressed in function of an arbitrary parameter u.
By considering the configuration spacetime R Q , the motion of S is described by a path of
points
(t (u ), q i (u ))
i = 1,..., n
which are simply denoted by

( z j (u ))

j = 0,1,..., n

with u [u1 , u 2 ] .
It is convenient to introduce the 2(n + 1) dimensional space of points and directions
associated with the timespace R Q and of which the elements are

( z j (u ), z j (u ))

j = 0,1,..., n

such that:
- for j = 0 :

z 0 (u ) = t (u ) ,

z 0 (u ) = t (u ) =

dt
du

390

Chapter 5

- for j 0 :

dq i
z (u ) = q (u ) =
du

z (u ) = q (u ) ,
j

i = 1,..., n.

From the above


z j = (t , q i ) ,

we deduce
z j = (1 ,

dq i du
q i
) = (1, ) = (1, q i ) .
du dt
t

The action integral


t2

S = L(t , q i , q i ) dt
t1

dq i du dt
L(t (u ), q (u ),
) du
dt du du

u2

u1

is simply written, by introducing L = L t , as follows:


S=

u2

u1

L( z j (u ), z j (u )) du

where the function L is assumed of class C 2 .


The (first) variation of S is
u2

L
z j ) du
j
z
z
L
L d
( j z j +
( z j )) du
j
du

z
z
L j d L
d L
( j z + ( j z j ) ( j ) z j ) du
du z
du z
z

S = (
u1

u2

u2

u1

u1

L
j

zj +

that is
u2

u1

S = (

L
d L
) z j du + [ j z j ]uu12 .
j
du z
z

Among the possible trajectories from (t1 , q1 ) to (t 2 , q 2 ) , these points corresponding to


respective values u1 and u 2 , we are going to prove that the trajectory fulfilling the
Lagranges and complementary equations is such that S = 0 .

First, we note the previous bracket vanishes.


Second, we have for j = 1,..., n :
L
z j

and for j = 0 :

d L
L
d ( Lt ' ) q i
L d
L z j

=
t

t
(
)
=
t

(
t
(
( ))
du z j
dt q i z j
q i
q i dt q i z j t
L d L
= t ( i
)=0
dt q i
q

Lagrangian Dynamics, Variational Principles


d L L
d
d
L
L dq i
L
(L + t ) =
(L + t i
)
t
t

=
du
du
t
t
q dt
z 0 du z 0 t

d
qi
d
L
L q i
L
L
( L + t i ( 2 )) =
)
t
t t (L i
du
dt
t
t
t
q t
q
L dL
L d
L
) = 0.
= t ( ( L i q i )) = t ( +
t dt
dt
q
t
=

We conclude that S = 0.

391

CHAPTER 6

HAMILTONIAN MECHANICS

In classical mechanics, the Hamiltonian theory, which dates from 19th century, does
not greatly simplify the differential equations of dynamic problems in comparison with the
Lagrangian version. On the contrary, the equations of motion can be often more easily
obtained from the Lagrangian method.
Nevertheless, the Hamiltonian formalism introduces the following advantage. The
Hamiltons equations may be simplified by transformations concerned by variables p i and q i
which are independent unlike the variables q i and q i in the Lagrangian approach.
Hamiltonian variables introduce a greater freedom, notably in celestial mechanics. Of
course, the Hamiltonian theory is really suited to quantum mechanics. Therefore, the area of
Hamiltonian dynamics is wider than the one of Lagrangian dynamics.
The Liouvilles theorem which states the invariance of the density in the phase space
shows a supplementary reason to consider the independent variables p i and q i of
Hamiltonian mechanics, this theorem being a foundation of statistical mechanics.
Finally, the Hamilton-Jacobi equation introduces a powerful way for solving
differential equations of motion.

1. N BODY PROBLEM AND CANONICAL EQUATIONS


The N-body problem consists in determining the trajectories of N particles interacting
in accordance with the gravitational law of Newton. Let us show the equations.
In an inertial frame of reference oxyz , we consider N particles p h of masses mh and
coordinates ( xh , yh , zh ).

393

394

Chapter 6

Each particle p h is attracted by the N 1 other particles pk with a force


N

fh =

mh mk

rhk2

k =1
k h

(6-1)

1hk

where

rhk
rhk

1hk =

( rhk = ph pk ).

We evidently have:

d 2 rh
dt

G mk

k =1
k h

rhk

( oph = rh ).

rhk3

The projections onto the coordinate axes lead to 3N differentials equations, namely:

mh

d 2 xh
dt

G mh m k ( x k x h )

k =1
k h

(( x k x h ) + ( y k y h ) 2 + ( z k z h ) 2 ) 3 2
2

and others in y and z.


The force function is written
U =

h, k
k h

mh mk
(( xk xh ) + ( yk yh ) 2 + ( zk zh ) 2 )1 2
2

(6-2)

where the sum is concerned with N(N-1) arrangements without repetition taken 2 by 2 (e.g. if
N=4, then 12, 13, 14; 21, 23, 24; 31, 32, 34; 41, 42, 43).
Since for any value l we have:
N
mk ml ( x k xl )
U
=G
2
2
2 32
xl
k =1 (( x k xl ) + ( y k y l ) + ( z k z l ) )
k l

then the 3N differential equations of motion are


mh xh =

PR1

U
,
x h

mh yh =

U
,
y h

mh zh =

z h

(6-3)

If the forces are conservative (that is derivable from a force function U ) then the
differential equation of dynamics can be put in a canonical form.

Proof. Consider a system of N particles subject to the gravitational law in R3 and let the
respective Cartesian coordinates denote
( x1 , y1 , z1 ) = ( x 1 , x 2 , x 3 ) ,
.
.
.
.
( x N , y N , z N ) = ( x 3 N 2 , x 3 N 1 , x 3 N ) .

395

Hamiltonian Mechanics

The kinetic energy of the system is


3N

T = 12 mi ( x i ) 2
i =1

where the common coefficient m1 = m 2 = m3 represents the mass of the first particle and so
on.
By letting
y i = mi x i

(no summation of course!)

we obtain
3N
( yi )2
T = 12
i =1 mi

T
x i

=0

T
y i

yi
mi

y i

= 0.

So, the system of motion equations (i = 1,,3N ):


U
x i

mi xi =

is equivalent to the following system of 6N equations


x i =

y i (T U )
=
,
mi
y i

y i =

(T U )
.
x i

If we introduce the function


F = T U

then the system of motion equations takes the canonical form:


x i =

F
y

y i =

F
x i

(i=1,,3N )

(6-4)

by recalling that the position of indices has no significance, it is only a notation to designate
3N coordinates.
Sometimes F is called a characteristic function and x i , y i are conjugate variables.
We have obtained a system of 6N differential equations of first order and there are 6N
unknowns: the positions and velocities of points. A priori 6N integrations are thus necessary,
but first integrals can be known.
By considering a problem with n degrees of freedom ( n = 3 N ), the characteristic function is
generally denoted by H and the system of canonical equations is written


x i =

H
y i

y i =

H
x i

(i=1,,n) .

(6-5)

396

Chapter 6
A system of motion equations is said to be canonical if there is a function H (t , x i , y i )
fulfilling (6-5).

Before introducing such a system of 2n canonical equations in the general Hamiltonian


approach of classical mechanics, let us briefly consider the problem of obtaining first
integrals.
PR2

A characteristic function H which does not explicitly depend on time is a first integral
of canonical equations.

Proof. From
dH
H
=
+
dt
t
=

H
+
t

H
x i x i +
i =1
n

H H

y i

y i

i =1

H H

( x i y i y i x i

i =1

we deduce
dH H
=
.
dt
t

So, if

H
= 0 , we have the first integral
t
H ( x i (t ), y i (t )) = c .

Remark. We mention that the case where H explicitly depends on t can be related to the one
where H does not explicitly depend on t, but the order of the corresponding system of
canonical equations is 2n+2 [e.g. Kovalevsky (1963)].

PR3

If a characteristic function does not explicitly depend on the independent variable t,


then the integration of the canonical system of order 2n is reduced to the one of a
system of order 2n-2, the time t being obtained from an integral.

Proof. We can view the system (6-5) as the following system of order 2n-1:
H
2

H
n

dx
y
=
,
1
H
dx

...

y 1

dx
y n
=
,
H
dx1
y 1

(6.6)
H

dy1
x1 ,
=

H
dx1
y1

with the first integral H ( x i , y i ) = c .

...

dy n
x n
=

H
dx1
y1

Hamiltonian Mechanics

397

The problem facing us is the integration of a system of order 2n-2; the solution will
express x 2 ,..., x n , y 1 ,... y n in function of x1 and c, time t being given by

t=

dx1
H
y

+ t0 .

(6-7)

Example. If n = 1 , the canonical system

x =

H
,
y

y =

H
x

is viewed as
x H dx + y H dy = 0

H ( x, y ) = c .

The integration of canonical equations is given by

t=

dx
+ t0 .
yH

2. CANONICAL EQUATIONS AND HAMILTONIAN

2.1

LEGENDRE TRANSFORMATION AND HAMILTONIAN

The Legendre transformation lets us establish the canonical equations of Hamilton


from the Lagrangess equations.
Given a function f : R R : x  f ( x) of class Cp ( p 2 ) and such that f ( x) 0 , we say:
D

The Legendre transformation of f ( with respect to x) is the function


L

f : R R : x L f ( x) = x f ( x) f ( x) .

(6-8a)

By introducing the inverse function of z = f ( x) :


g : R R : z  x = g ( z) ,

the Legendre transformation of f is viewed as


L

f : R R : z  L f ( z ) = z g ( z ) f ( g ( z )) .

(6-8b)

398

Chapter 6

Remark. Given the assumptions, the second derivative ( L f )( z ) does not vanish since we
have:
d

dz

f ( z) = g ( z) + z

dg
dz

df dg
dx dz

= g ( z)

which really implies:


( L f )( z ) = g ( z ) =

1
0
f ( x)

In addition, letting L f ( z ) = F ( z ) , we have (involution!):


LL

f ( z ) = LF ( z ) = z

dF
dz

( z ) F ( z ) = z g ( z ) ( z g ( z ) f ( g ( z ))

= f ( g ( z )).

Example. What is the Legendre transformation of the function f ( x) =

x2 ?

Answer. We note that


d
f ( x) = mx
dx

d2
f ( x) = m 0 .
dx 2

Next, the inverse function of z = f (x) is


x = g ( z) =

z
m

and thus
L

f ( z) = z

z m z 2 z2
( ) =
.
m 2 m
2m

Generalization. The definition of the Legendre transformation becomes immediately


widespread to functions of vector variables x = ( x1 ,..., x r ) and y = ( y 1 ,..., y s ) .

Thus we consider a function f (x,y) of class C p ( p 2 ) for x such that


det(

2 f
x i y j

) 0.

The system of r equations


zk =

f
( x, y)
x k

can be locally solved with respect to the r variables xi, that is


xi = i ( z, y )

and we express:

i = 1,,r

399

Hamiltonian Mechanics
D

The Legendre transformation of f (x,y), with respect to x, is the real-valued function


from R r R s :
f
Lx
(6-9)
f = i i f ,
x
that is explicitly
Lx

f ( z, y ) =

f i
( z , y ) f ( x ( z , y ), y ) .
x i

Before introducing the canonical equations, we are going to define the Hamiltonian function
(or Hamiltonian) as a Legendre transformation of Lagrangian.
We recall that the function
L : R R n R n R : (t , q i , q i )  L(t , q i , q i )

(6-10)

is the Lagrangian T-V (at lest of class C2 ) of a material system with n degrees of freedom.
We also recall that the generalized momentum or canonically conjugate momentum to qi is
pi =

L
T
= i .
i
q
q

(611)

2L
We assume that the condition det( i j ) 0 is fulfilled.
q q

If the various q i are likened to the previous xi, if (t , q1 ,..., q n ) is likened to the previous y and
if f is taken to be the Lagrangian, then we can set:
D  The Hamiltonian is the Legendre transformation of L(t , q i , q i ) with respect to
q = (q 1 ,..., q n ) , that is

H = pi q i L

i = 1,,n

(6-12)

where the various q i are expressed in function of 2n + 1 variables t, pj, qk.


Explicitly:
H (t , p, q ) =

L q

L(t , p, q ) = pi q i (t , p, q ) L(t , q, q (t , p, q )) .

Example 1. Write down the Hamiltonian for a free particle of mass m which moves, in R3,
under the action of forces derivable from a potential V.

Answer. From the Lagrangian


L=

m
2

( x 2 + y 2 + z 2 ) V ( x, y, z )

we deduce the following generalized momenta

400

Chapter 6
px =

L
= mx ,
x

p y = my ,

p z = mz .

In this example, we mention that the generalized momenta are the linear momenta.
Since
py
p
p
x = x ,
y =
,
z = z
m
m
m
we obtain
2
p x2 p y p z2
H=
+
+
L
m
m
m

2m

( p x2 + p 2y + p z2 ) + V ( x, y, z ).

Example 2. Determine the Hamiltonian of a particle of mass m moving in a central field.

Answer. From the Lagrangian


L=

m
2

(r 2 + r 2 2 ) V (r ) ,

we deduce the generalized momenta


pr =

L
= mr ,
r

p =

L
= mr 2 .


 =

From
r =

pr
,
m

mr 2

we obtain

H=

p2
p r2
+ 2 + V (r ) .
2m 2mr

Example 3. Find the Hamiltonian of a free particle of mass m which moves in a frame of axes
oXYZ rotating relative to one another of axes oxyz about the common axis oz = oZ with a
constant angular velocity = 1 . The force is derivable from a potential V ( X , Y , Z ) .

Answer. For this rheonomic problem, from


x cos t

y = sin t
z 0

sin t
cos t
0

0
1

X

Y
Z

we deduce
T =

and thus

m
2

( x 2 + y 2 + z 2 ) =

m  2 2  2
m
( X + Y + Z ) + m( XY XY ) + ( X 2 + Y 2 )
2
2

401

Hamiltonian Mechanics
pX =

L
= mX mY ,
X

L
= mY + mX ,

Y
L
pZ =
= mZ
Z

pY =

p
X = X + Y ,
m
p
Y = Y X ,
m
p
Z = Z
m

Therefore, the Hamiltonian is


H=

2.2

1
2m

( p 2X + pY2 + p Z2 ) + (Y p X X pY ) + V ( X , Y , Z ) .

CANONICAL EQUATIONS

Let us see that the Legendre transformation converts the Lagrange equations into
canonical equations of the Hamiltonian formalism.
PR4

A system of 2n first order differential equations called the system of canonical


equations of Hamilton and following from the system of n Lagrange equations is
written:
q i =

H
,
pi

p i =

H
.
q i

(6-13)

Proof. Since
L i L i L
dt
dq i dq
t
q
q i
L
L
dt
= q i dpi i dq i
t
q

dH = pi dq i + q i dpi

and
dH =

H
H i H
dt ,
dpi +
dq +
i
t
pi
q

we deduce:
q i =

H
,
pi

H
L
= i ,
i
q
q

L
H
.
=
t
t

The system of n Lagranges equations:


p i

L
=0
q i

leads to the system of 2n canonical equations:


q i =

H
,
pi

p i =

H
q i

describing the evolution of the material system. These equations have the canonical form of
equations (6-3) where F is the Hamiltonian H.

402

Chapter 6

Remark. The equations


L
H
=
t
t

and
H H
dH H
H
=
p i + i q i +
=
t
t
dt
q
pi

introduce the complementary equation:


dH L
=0
+
dt
t

(6-14)

which takes the following form in Lagrangian formalism:


d L
=0
L +
t
dt

(6-15)

where L = pi q i L is the well-known adjoint Lagrangian.


Of course, L is H if we express the various q i in function of generalized momenta.
PR5

The canonical equations follow from the variational principle of Hamilton (where
L = pi q i H ).

Proof. By recalling (5.41), Hamiltons principle is written in this new context as


t2

( pi q i H )dt = 0 .

(6-16)

t1

We can consider again the variational problem developed in Section 2 of Chapter 5 with the
function

f = pi q i H .
The condition of extremum S = S (0) = 0 is here written
f
f
+ q i i + p i
1
pi
q
t2
H
= ( pi q i pi
qi
t1
pi
t2

( pi

f
f
+ q i i ) dt
p i
q
H
+ q i pi ) dt = 0 .
i
q

But we know that the last term is


t2

pi q i dt =

t2

pi
t2

t2
d i
q dt = [ pi q i ]tt12 p i qi dt
t1
dt

= p i q i dt
t1

since q i (t1 ) = q i (t2 ) = 0 .


Thus Hamiltons principle is written:

(sum over i)

403

Hamiltonian Mechanics
t2

( (q i

H
H
)pi ( p i + i )q i ) dt = 0 .
pi
q

(6-17)

Given arbitrary variations q i and pi of independent variables qi and pi, we deduce the
canonical equations
q i =

H
,
pi

p i =

H
q i

Remark. Once more, the independence of variables qi and pi let us conclude. Of course, as
soon as the motion is determined, these coordinates are no longer independent since, from qi,
L
dq i (t )
and pi = i (t , q, q ) .
we obtain q i (t ) =
dt
q
Example. Find the canonical equations for a particle moving in a central force field.

Answer. From

H=

p2
p r2
+ 2 + V (r )
2m 2mr

we deduce
r =

p
H
= r ,
pr
m

p r =

p 3 dV
H
,
= 3
r mr
dr

 =

H
p
= 2
p mr

p =

H
= 0.

Phase space
We know that any material system determines a point (q1 ,..., q n ) in the configuration
space. But the evolution of a material system is specified in addition by the generalized
velocities q 1 ,..., q n and thus by the generalized momenta.
Given an open U R n , we note that pairs (p,q) are points of a vector space U R n and we
say:
D

The space of 2n-tuples ( pi , q i ) of independent conjugate variables is called the phase


space associated with a material system.

We denote this 2n-dimensional space by P and thus


( p1 ,..., p n , q1 ,..., q n ) = ( p, q) P .
The motion of a given material system is represented by a phase point (p,q) which describes a
curve in the phase space in accordance with the corresponding Hamiltons canonical
equations.

404

2.3

Chapter 6
FIRST INTEGRALS AND CYCLIC COORDINATES

We have already defined the first integral notion in the Lagrangian context and it is
immediately transposable to the Hamiltonian formalism.
PR6

The Hamiltonian of a scleronomic material system is constant during the motion.

Proof. By considering the complementary equation, we have


0=

L
dH
=
t
dt

H =C.

So, in the case of scleronomic and conservative systems (with potential V ), PR6 of Chapter 5
becomes:
PR7

Every scleronomic and conservative system has the Hamiltonian as a first integral
called the Jacobi integral, which is nothing else than the mechanical energy H = E .

The proof is like the one of the recalled proposition; namely, from the Euler theorem about
the homogeneous functions, we deduce
H = pi q i L =

T i
q L = 2T L = T + V .
q i

(6-18)

We recall the following


D

A generalized coordinate qk is cyclic if it does not explicitly occur in the Lagrangian or


Hamiltonian:
L
q

=0

H
q k

= 0.

In this case, we find again PR7 of Chapter5; that is, the generalized momentum conjugate to
qk is a first integral:
H
q k

=0

p k = 0

pk = c .

Relevance of cyclic coordinates


If a coordinate (e.g. qn) is cyclic, then the conjugate momentum is a first integral
( p n = c ). So the Hamiltonian is a function of other qi, of other pi, of an arbitrary constant c
(determined by the initial conditions) and of t, namely:

H (t , p1 ,..., p n1 , q1 ,..., q n 1 , c) .
So the problem comprises n 1 generalized coordinates. Its solution will let us determine the
H
.
cyclic coordinate by integrating q n =
c

405

Hamiltonian Mechanics

Remark 1. To integrate the canonical equations is to search the independent first integrals. If
all these were known, the problem (and thus its solution) would be reduced to a system 2n
finite equations.
Remark 2. The first integrals are deduced from the canonical equations. So, in the example
of a central force field, we have
p =

which implies that

H
=0

p = mr 2

is constant during the motion. We find again a well-known result since the coordinate is
cyclic.
Remark 3. It turns out to be necessary to choose a coordinate system that gives the
maximum of cyclic coordinates, this choice making the most of symmetries of the problem.
In the previous example, no Cartesian coordinate would be cyclic.
The Rouths method.

If a cyclic coordinate exists, for instance qn, we know that the Hamiltonian
H (t , p1 ,..., p n1 , q1 ,..., q n 1 , c)
depends on an integration constant.
In this context, the Rouths method consists in combining at once the Lagrangian and
Hamiltonian formalisms.
Given r cyclic coordinates q 1 ,..., q r , we say:
D

The Routhian is the function R defined by


r

R(t , q1,..., q n , p1,..., pr , q r +1,..., q n ) = pi q i L(t , q, q ) .


i =1

The differential
r

dR = q i dpi +
i =1

q i dpi
i =1

pi dq i L dt Li dqi Li dq i


t
i =1
i =1 q
i =1 q
n

pi dq i Li dqi L dt

t
i =1 q
i = r +1

implies
i {1,,r}:
i { r + 1 ,,n}:

R
= q i ,
p i
L
R
= i
i
q
q

q i
R
q

= p i
=

L
q i

(6-19)

406

Chapter 6

The first collection of 2r equations is the one of canonical equations for the r cyclic
coordinates with the Hamiltonian R, the second collection of equations fulfills the n r
Lagranges equations:
d R
R
( i) i =0
dt q
q

i = r + 1,..., n .

But, for the cyclic coordinates, we have:


i {1,,r}:

L
q

=0

R
q i

=0

and the corresponding momenta pi are arbitrary constants ci to be determined from the initial
conditions.
So, the n r Lagranges equations are solved with the Routhian
R (t , q r +1 ,..., q n , q r +1 ,..., q n , c1 ,..., c r ) .

2.4

LIOUVILLES THEOREM IN STATISTICAL MECHANICS

We know that any point of the phase space, called phase point, determines a possible
state of a given material system. At the initial time, the state of the system is determined by a
i
2n-tuple (pi,q ) in the phase space and the motion of the system is followed by a representative
point which describes a unique path in the phase space according to the 2n Hamiltons
equations.
Of course, if the number of degrees of freedom of the material system is very large and thus
the number of canonical equations too, then obtaining the (unique) solution of the system of
these equations can be impossible in practice. In addition, in the case of systems containing a
great number of particles, it will be impossible to know the initial conditions of each of them.
In order to overcome such difficulties, we are going to introduce a fundamental theorem of
statistical mechanics due to Liouville.
Whereas a well-determined phase point cannot be specified to represent the state of a system,
we consider a set of neighboring phase points. These phase points represent neighboring
states of the one of the material system and determine a cloud of possible states of the given
system. So, we imagine a collection of phase points such that any of which could potentially
represent the material system; in other words, we imagine a great number of systems
corresponding to a given material system but with slightly different initial conditions.

Fig. 83

407

Hamiltonian Mechanics

The motion of each phase point represents the (independent) motion of a particular system.
Let be the phase volume element occupied by the neighboring phase points at the initial
time t0 and let denote the volume element occupied by the previous phase points at a later
date t 0 + dt.
We note that the number of representative phase points of a material system is the same in
and in since any phase point initially within can never leave this volume. Indeed, if a
representative phase point of a possible system escaped at a given time, that would mean this
point would occupy the same position as a phase point belonging to the boundary of the
volume; but, since the (unique) motion is exactly determined from a given position of the
phase space, then two corresponding material systems would move in the phase space along
the same path.
We emphasize that any phase path representing the (independent) motion of any system of ,
that is a particular solution of the Hamiltons equations, cannot intersect another and the
number of representative phase points is invariable! Now, we can prove the following
Liouvilles theorem.
PR8

The density of representative points in phase space corresponding to the motion of a


given material system is invariable during the motion.

Proof. According to the preliminaries, we must prove that the volumes and are equal,
which will imply that the number of representative phase points per unit volume remains
constant during the motion.
During a time variation dt all the phase points of move and their coordinates vary from pi,
qi to the following

pi = pi + dpi ,

q i = q i + dq i .

(6-20)

The relation between corresponding volumes and is obviously


=
where J =

( p' , q ' )

( p, q )

( p' , q ' )
is the Jacobian of the coordinate transformation defined by (6-20).
( p, q )

From

p'i = pi + p i dt ,

q i = q i + q i dt

pi
p
= ij + i dt ,
p j
p j

pi

we deduce

q i q i
dt ,
=
p j p j
and therefore the Jacobian is expressed as

p i
q j
q i
q j

dt

= ij +

q i
q j

dt

408

Chapter 6

1+

p 1
dt
p1

p 1
dt
p 2

...

p 1
dt
p n

...

p 1
q n

dt

.
J=

.
.
q n
dt
p1

q n
dt
p 2

...

q n
q n
dt ... 1 + n dt
p n
q

By limiting the expansion of J to the first order in dt, we obtain:


p i q i
+ i ) = 0,
J =1 + (
q
i =1 p i
n

but from
n

(
i =1

p i q i
+
)=0
pi q i

we deduce that J = 1 and we have really obtained the expected equality


' = .

From this result, we are going to prove that the density (t , p, q ) of points in the phase space,
such that the number of systems (whose representative points lie within ) is equal to ,
remains constant during the motion.
Indeed, by expressing that this number is constant:
d
( ) = 0
dt

and knowing that


d
= 0,
dt

we deduce:

and thus

d
=0
dt
d
= 0,
dt

which means that the density of possible systems or the number of phase points per unit phase
volume is invariable during the motion:
d

=
+
p i + i q i = 0 .
t pi
dt
q

(6-21)

Remark. In differential geometry, the phase volume is denoted by dp dq. Since the
number of representative points of (moving in the phase space) is invariable, that is
d
d
d
( dp dq ) = 0 , and since
(dp dq ) = 0 , we deduce
= 0.
dt
dt
dt

Hamiltonian Mechanics

409

3. CANONICAL TRANSFORMATIONS
The choice of generalized coordinates and generalized momenta is not unique; if we
know that the Lagrange equations are invariable under changes of generalized coordinates, on
the contrary we will have to search for transformations preserving the canonical form of
Hamiltons equations. Transformations called canonical have this property and such
particular transformations play a fundamental role in Hamiltonian mechanics, in symplectic
geometry and in celestial mechanics for instance. Such transformations can be very helpful
insofar as new coordinates are cyclic. We are going to explain that.

3.1

LAGRANGE AND POISSON BRACKETS


We have already defined the phase space P and we say:

The phase spacetime or state space is


RP .

3.1.1

Preservation of Canonical Form and Poisson Bracket


First, we note the following

PR9

Any diffeomorphism1 on the state space does not necessarily preserve the form of
Hamiltons canonical equations.

Proof. Let us consider the canonical system


q i =

H
,
pi

p i =

H
q i

and the following diffeomorphism

Q I = Q I (t , pi , q i ) ,

PI = PI (t , pi , q i )

I , i = 1,..., n.

In the new coordinates, the Hamiltonian H is written as follows:


K (t , P , Q ) = H (t , p(t , P , Q ), q (t , P , Q )) .

Before continuing, let us introduce the definition of the Poisson bracket.


Given an open I of R, opens and of Rn, we consider two functions at least of class C1,
namely f and g: I R respectively defined by f (t , p, q ) and g (t , p, q ) .
D  The Poisson bracket 2 of functions f and g is the function defined by


{f,g} = (
i

1
2

f g f g
).

q i pi pi q i

Bijection which is differentiable as well as its inverse (both of class C2 for instance).
For instance, the Liouvilles theorem is conveniently expressed as
d
=
+ {,H} = 0
dt
t

(6-22)

410

Chapter 6

Under the diffeomorphism the equations of Hamilton take the following form:
Q I
Q I H Q I H
Q I =
+ ( i

)
t
q p i pi q i
i
=

Q I
Q I K Q J K PJ
Q I K Q J K PJ
+
+

(
)
i p ( Q J q i + P q i )
t
PJ p i
q i Q J pi
i
i
J

Q I
K
K
+ { QI ,QJ }
+ { Q I , PJ }

J
PJ
t
Q

In the same manner we obtain:


P
K
K
+ { PI , PJ }
.
PI = I + { PI , Q J }
J
t
Q
PJ

However the canonical form of Hamiltons equations is preserved after transformation iff
there is a function H (t , PI , Q I ) such that
H
Q I =
,
PI

H
PI = I ,
Q

that is not always the case, which proves the proposition.


Now, the question is to know under what conditions the canonical form is preserved; namely:
The diffeomorphism defined by

PI = PI (t , p, q) ,

Q I = Q I (t , p, q )

preserves the canonical form of Hamiltons equations if


{ Q I , Q J }={ PI , PJ } = 0
{ Q I , PJ } = JI

( { PI , Q J } = IJ )

and if there is a function F (t , PI , Q I ) such that


QI =

F
,
PI

PI =

F
.
Q I

Indeed, we have:
2F
K
F
+
Q I =
=
(
+ K),
t PI PI PI t

2F
K
F
PI =

= I(
+ K) ,
I
I
t Q
Q
Q t
with the following transformed Hamiltonian
H =

F
+K.
t

Hamiltonian Mechanics
3.1.2

411

Poisson Bracket and Symplectic Matrix

Among other things, the Poisson brackets let us:


verify if a function f (t , p, q ) is a first integral,
find new first integrals from known first integrals,
express the motion equations under a very simple form, and so on.

Let us denote the coordinates of any point x of the phase space by


x A = ( pi , q i )
where A{1,,2n} and i{1,,n}.
We introduce the following (2n 2n) matrix the importance of which will be later underlined.
D  The matrix
0
J = ( AB ) =
I

I
,
0

(6-23)

where I is the ( n n ) unit matrix, is called the symplectic matrix.


The inverse matrix is

J 1 = AB

such that

AB BC = AC .
Of course, we have
J 2 = I
and
0 I
.
J 1 = t J = J =
I
0

So, the symplectic matrix is antisymmetric and of determinant +1.


The Poisson bracket of functions f and g is expressed as


{ f , g } = AB

f g
.
x A x B

(6-24)

In matrix notation, this is written:

{ f , g } = f

g

p
g

q

412

Chapter 6
t

f

p
=
f

q

g

p .
g

q

PR10 The 2n Hamiltons equations are denoted by


H
x B

x A = AB

(6-25)

with A, B {1,...,2n } .
Proof. For the n first indices these equations are

p i = iB

= i n+ j

x B
and for the n last indices they are
q i = n+i B

H
x

H
x n + j

= n+i j

H
x

H
q i

H
.
p i

Canonical equations and Poisson bracket


The expression of the Poisson bracket of functions f (t , p, q ) and H (t , p, q ) is

{ f , H } = ( f i
n

i =1

H f H
)

q p i p i q i

that is


{ f , H } = AB

f H
.
x A x B

(6-26)

PR11 The canonical equations of Hamilton have the following elegant expression:

x A = { x A , H }

(6-27)

Proof. We have

x A H
H
= BC BA C
x ,H =
B
C
x x
x

H
= AC C
x
A
= x .
A

BC

So, for A = 1,..., n and A = n + 1,...,2n respectively, we have:

p i = { pi , H },

q i = { q i , H }.

(6-28)

413

Hamiltonian Mechanics
First integral existence
PR12 The existence condition of a first integral f of canonical equations of Hamilton is
f
+ { f , H }=0.
t

(6-29)

Proof. This is obvious since


n
df f
f dp i f dq i
)
+
+ (
=
dt t i =1 pi dt q i dt
f
+ {f , H } = 0.
=
t

Lie algebra
First, we recall that, given a commutative field K, an algebra on K or K-algebra is a
vector space E on K provided with a bilinear mapping *: E E K .
In other words, we say:
D

A K-algebra is a vector space K , E ,+ provided with an (inner) law * such that:


x, y, z E : ( x + y ) z = ( x z ) + ( y z ) ,
x , y, z E : x ( y + z ) = ( x y ) + ( x z ) ,
k1, k2 K , x , y E : (k1 x ) (k2 y ) = k1k2 ( x y ) .

In addition, we say:
D

A Lie algebra is an algebra for which the inner law is anticommutative and satisfies
the Jacobis identity, that is:
x , y E : x y = y x ,
x , y, z E : ( x y ) z + ( y z ) x + ( z x ) y = 0.

Let C ( R P ) denote the space of real-valued functions on a state space.


PR13 The space C ( R P ) together with the Poisson bracket forms a Lie algebra.

Proof. From the Poisson bracket definition, we have immediately,

k1, k2 R, f , g , h C ( R P ) :
(i) The bilinearity:

{ f , g + h} = { f , g} + { f , h} ,
{ f + g , h} = { f , h} + {g , h} ,
{k1 f , k 2 g} = k1k 2 { f , g} .

(6 30)

414

Chapter 6

(ii) The anticommutativity:

{ f , g} = {g , f } .
(iii) Jacobis identity:

{{ f , g}, h} + {{g , h}, f } + {{h, f }, g} = 0 .

(6-30)

Let us prove this last with the help of AB :

f g h

g h f
( CD C
) B + AB A ( CD C
)
A
D
x
x x x
x
x x D x B

h f g
+ AB A ( CD C
)
x
x x D x B

AB

2 f
g
f
2g
h
2 g h
g 2 h
f
AB CD

)
(
) B
+
+
+
A
C
D
C
A
D
B
A
C
D
C
A
D
x x x
x x x x
x x x
x x x x
2
2
h f
h f
g
) B .
+ AB CD ( A C
+ C
D
A
D
x x x
x x x x

= AB CD (

The second and third terms cancel and so on. Indeed, by changing the names of summation
indexes A D, B C , C A, D B in the third term, this last becomes

DC AB

2 g h f
,
x D x A x B x C

that is the second term opposite.


In the same manner, the fourth term cancels the fifth and the sixth term cancels the first.
The two additional properties:
{ f , gh } = g { f , h } + h { f , g },

(6-31)

{ f , g } = { f , g } + { f , g }
t
t
t

(6-32)

are immediately proved from the Lagrange bracket definition.

Poisson theorem

PR14 If f and g are first integrals of Hamiltons equations, then their Poisson bracket is a
first integral.
Proof. From
and
we deduce:

t f + {f , H} = 0 ,

t g + {g , H } = 0

{H , { f , g}} + { f , {g , H }} + {g , {H , f }} = 0 ,
{H , { f , g}} { f , t g} + {g , t f } = 0

415

Hamiltonian Mechanics

and thus

{ f , g } + {{ f , g }, H } = 0 .
t

From this theorem of Poisson, we can prove the following


PR15 Given a conservative system ( H = E ), the existence of a first integral f (t , p, q )
implies that the successive derivatives

i f
are first integrals.
t i

Proof. The theorem of Poisson lets us assert that


integral iff
t f + {f , H} = 0

{ f , H } is

a first integral. But f is a first

and thus t f is a first integral.


Then, Poissons theorem means that { t f , H } is a first integral. But t f is a first integral iff
2 f
+ { t f , H } = 0 ,
t 2
2 f
thus we conclude
is a first integral, and so on.
t 2
D

Any two functions f and g are in involution or involutive if their Poisson bracket is
zero.

PR16 If a function f is involutive with other functions g and h, then it is involutive with their
Poisson bracket {g , h}.

Proof. Since we have { f , g } = { f , h} = 0 , the identity of Jacobi implies

{ f , {g , h}} = {g , {h, f }} {h, { f , g}} = 0 .


3.1.3

Lagrange and Poisson Brackets

Given a diffeomorphism on the phase space:

Q I = Q I ( pi , q i ) ,

PI = PI ( pi , q i ) ,

then for any P and Q of { P1 ,..., Pn , Q1 ,..., Q n } we say:


D


The Lagrange bracket of coordinates P and Q is the real-valued function:


n

( P, Q ) = (
i =1

pi q i q i pi
).

P Q P Q

(6-33a)

416

Chapter 6

The diffeomorphism being denoted by y A = y A ( x B ) with A, B {1,...,2n}, the previous


definition is expressed as follows:
D

The Lagrange bracket of any two coordinates yC and yD is the real-valued function
( y C , y D ) = AB

x A x B

y C y D

(6-33b)

PR17 There is the following relation between Poisson and Lagrange brackets:
2n

(y

, y D ) y D , y B = cB .

, y D ) {y D , y B } = AE

D =1

(6-34)

Proof. We have:
2n

(y
D =1

= AE RE RS

3.2

x A x E RS y D y B

y C y D
x R x S

A
B
A
B
x A y B
ES x y
S x y
=
=

= CB .
AE
A
C
S
C
S
C
S
y x
y x
y x

CANONICAL TRANSFORMATION

We are going to consider canonical transformations (which enrich Hamiltonian


mechanics); for instance, such transformations of generalized coordinates and momenta
preserve the canonical structure of the motion equations.
Let us consider a neighborhood in a configuration space and the corresponding domain in the
phase space.
D  A diffeomorphism

Pi = Pi ( p, q ) ,

Q i = Q i ( p, q )

is a canonical transformation on the phase space if the differential


pi dq i Pi dQ i
is exact;
in other words, if there is a real-valued function S on the phase space such that


pi dq i Pi dQ i = dS

(6-35)

dpi dq i = dPi dQ i .

(6-36)

or if


417

Hamiltonian mechanics
Example. Given
Pi = f (q i ) cos pi ,

i = 1,..., n ,

Q i = f (q i ) sin pi

let us specify f (q i ) such that this transformation be canonical.


Answer. Since
f
f
cos pi dq i f sin pi dpi ) ( i sin pi dq i + f cos pi dpi )
i
q
q
f
= i f dpi dq i ,
q

dPi dQ i = (

the transformation is canonical iff i { 1,..., n }:

f
f =1
q i

1
2

( f )2
= 1
q i

that is

( f (q i )) 2 = 2q i + c i

( ci R ) .

3.2.1 Canonical Transformations and Brackets


PR18 A diffeomorphism
Pi = Pi ( p, q ) ,

Q i = Q i ( p, q )

is a canonical transformation on the phase space iff


( Pi , Pj ) = 0 ,

Proof. We consider

(Q i , Q j ) = 0 ,

( Pi , Q j ) = i j .

(6-37)

( P , Q )
0.
( p, q)

The differential form p r dq r Ps dQ s is written in function of variables Pi and Qi as follows:


q r
q r
pr ( P , Q)
dPs + ( p r ( P , Q )
Ps ) dQ s
s
Ps
Q
(summation over r and s).

It is exact iff
Ps
p r q r
p r q r
2qr
2q r
+
=
+

p
p
r
r
Pk Q s
Q s Pk
Pk Q s
Q s Pk Pk

P
and, since s = sk , it is exact iff
Pk
p r q r q r p r

= ( Pk , Q s ) = sk ;
s
s
Pk Q
Pk Q

()
r

418

Chapter 6

in addition, since

Ps
Q

Pm

= 0 , the condition

Q s

p r q r
Q

Ps
Q

p r q r

Q Q
s

Pm
Q s

becomes
p r q r
Q Q
s

q r p r
Q Q
s

= (Q s , Q m ) = 0 ;

finally, the condition


p r q r
p r q r
2q r
2q r
+ pr
=
+ pr
Pm Ps
Ps Pm
Ps Pm
Pm Ps

is
p r q r q r p r

= ( Pm , Ps ) = 0 .
Pm Ps Pm Ps

By recalling the zero square property of the exterior differentiation ( d  d = 0 ), we conclude


that a canonical transformation on the phase space is a differentiable mapping which
preserves the 2-form

= dpi dq i = dPi dQ i .
Important remark.

We denote a general 2-form 2 (P ) as follows:

= 12 AB dx A dx B .

(6-38)

In the following explicit expression

= 12 ( ij dx i dx j + i n+ j dx i dx n + j + n+i j dx n+i dx j + n +i n + j dx n +i dx n+ j ) ,
we notice that the various ij are zero since there is no term of type dpi dp j , the various

n+i n+ j are zero since there is no term of type dq i dq j and the only nonzero terms are
i n+i = n+i i = 1 .
Of course, the symplectic matrix J = ( AB ) appears.

The previous skew-symmetric invariant bilinear form plays a fundamental role in symplectic
geometry, this form being necessary to study the natural symplectic structure of the phase
space. Such a study is outside the field of this book, but it is introduced, for instance, in
Talpaert [2000].
PR19 A diffeomorphism on P is a canonical transformation
iff the matrix associated to the Lagrange bracket is the symplectic matrix J,
iff the matrix associated to the Poisson bracket is J 1 .

419

Hamiltonian mechanics

Proof. A diffeomorphism y A = y A ( x B ) defines a canonical transformation


iff
x A x B
CD dy C dy D = AB dx A dx B = AB C D dy C dy D
y y
= ( y C , y D ) dy C dy D

(sum over C and D )

iff
( y C , y D ) = CD

iff

{y

, y D } = CD

[since (6-34)].

Remark. On the phase space, the existence criterion of a canonical transformation is

( PI , PJ ) = 0 ,

(Q I , Q J ) = 0 ,

( PI , Q J ) = IJ

{PI , PJ } = 0 ,

{Q , Q } = 0 ,

{Q

, PI } = .

(6-39)

J
I

PR20 A transformation is canonical iff it preserves the Poisson bracket.


Proof. Given a transformation y A = y A ( x B ) , we have:
g y A y B
f g
CD f
=
{ f , g }x =
x C x D
y A y B x C x D
f g
= { y A, yB } A B ,
y y
CD

but
{ f , g }y = AB

f g
y A y B

and thus
{ f , g }y = { f , g }x

AB = { y A , y B }

that is iff the transformation is canonical.

3.2.2

Canonical Transformations and Generating Functions

Let us consider a neighborhood in a configuration space and the corresponding domain


in the state space R P .
D

A diffeomorphism defined by
Pi = Pi (t , p, q) ,

Q i = Q i (t , p, q ) ,

t =t

is a canonical transformation on the state space R P if the differential


( pi dq i Hdt ) ( Pi dQ i Hdt )
is exact;

420

Chapter 6

in other words, if there is a real-valued function S on the state space such that
pi dq i Hdt = Pi dQ i Hdt + dS

(6-40)

dpi dq i dH dt = dPi dQ i dH dt .

(6-41)

or if

By recalling PR5, we know that the variational principle holds for two collections of
canonical variables {pi,qi} and {Pi,Qi} at once, that is
t2

( pi q i H ) dt = 0 ,
t1

( PI Q I H )dt = 0 .
t2

t1

By referring to PR13 of Chapter 5, it is certain that the two previous integrands do not differ
dS
of an arbitrary function S which does not depend on
from more than the total derivative
dt
variables q i and Q I too.
So, the variation of
t2

t2
t 2 dS
( pi q i H )dt ( PI Q I H )dt =
dt = S ( 2) S (1)
t1
t1 dt

vanishes at limits (1) and (2) whatever the described function S.


D

The previous function S (t , pi , q i , PI , Q I ) such that


pi dq i PI dQ I = ( H H )dt + dS

is called the generating function of the canonical transformation.


The generating function S is a priori dependent on 4n+1 variables. However 2n+1 variables
are independent in view of the following 2n relations:
PI = PI (t , p, q) ,

Q I = Q I (t , p, q ) .

Consequently, any generating function is one of the following types:


S1 (t , q, Q ) , S 2 (t , P , q) , S 3 (t , p, Q ) , S 4 (t , p, P ) .

(i)

Generating function S1 (t , q, Q ) .

(qi , Q i )
0 in such a way that qi and Q i are independent coordinates.
We suppose det
i
( pi , q )
We have:
S
S
S
pi dq i H dt = Pi dQ i H dt + 1 dt + 1i dq i + 1i dQ i
t
q
Q

421

Hamiltonian mechanics

pi =

S1
,
q i

S1
,
Q i

Pi =

H =

S1
+H .
t

(6-42)

2S
These relations define the canonical transformation (non singular if det i 1 j 0 ).
Q q
Example. The generating function

S1 = q i Q i
i

generates the following canonical transformation


pi = Q i ,

( H =H ).

Pi = q i

This transformation is assuredly canonical because dp i dq i = dPi dQ i is fulfilled.


By considering the generating function S1 , let us prove the following
PR21 Canonical transformations preserve the canonical form of Hamiltons equations.

Proof. Given the canonical equations


p i =

H
,
q i

q i =

H
pi

and the canonical transformation


Pi = Pi (t , p, q ) ,

Q i = Q i (t , p, q ) ,

we have:
P
P
P H PI H
PI = t PI + I p i + Ij q j = t PI I
+
pi
pi q i q j p j
q
=

2S
+ {PI , H }
Q I t

[ since (6-42) ] .

But, the Poisson brackets are invariant under the canonical transformations, therefore we
have:
{PI , H } = HI
Q
and thus
2 S1
H
PI =

I
Q t Q I
H
=
[ since (6-42) ].
Q I
The invariance of the other Hamilton equations is proved in the same way.

422

(ii)

Chapter 6
Generating function S 2 (t , P , q)

The generating function S2 is defined from the Legendre transformation of S1 with


respect to Q:
L

S2 (t , P , q) = Q S1 (t , q, Q ) = (

S1 i
Q S1 (t , q, Q ))
Q i

= PiQi + S1 (t , q, Q ).
From

pi dq i Hdt = Pi dQ i H dt + dS1 (t , q, Q )
we deduce
dS 2 (t , P , q) = pi dq i + Q i dPi + ( H H ) dt .
and thus
pi =

S 2
,
q i

Qi =

S 2
,
Pi

H =

S 2
+H
t

(6-43)

with the condition


2S
det i 2
q Pj

(iii)

0.

Generating function S 3 (t , p, Q )

The generating function S3 is defined from the Legendre transformation of S1 with


respect to q:
S
L
S 3 (t , p, Q ) = q S1 (t , q, Q ) = ( 1i q i S1 (t , q, Q ))
q
= pi q i + S1 (t , q, Q ) .
Differentiation leads to
Pi =

S 3
,
Q i

qi =

S 3
,
pi

H =H+

S3
t

(6-44)

with

2 S3
0.
det
j
p
Q

(iv)

Generating function S 4 (t , p, P )

The generating function S4 is defined from the Legendre transformation of S2 with


respect to q:

423

Hamiltonian mechanics
L

S 4 (t , p, P ) = q S 2 (t , P , q) = (

S 2
q

q i S 2 (t , P , q))

= S1 (t , q, Q ) + Pi Q pi q .
i

Differentiation leads to
qi =

S 4
,
pi

S 4
,
Pi

Qi =

H =

S 4
+H.
t

(6-45)

with
2S4
det
pi Pj

0.

Real symplectic group

We are going to see that the generating functions for canonical transformations are
obviously related and we are going to consider an essential matrix of second derivatives of
generating functions.
Given a diffeomorphism on the phase space defined by
y A = y A (x B )

A, B {1,...,2n}

Pi = Pi ( p, q ) ,

Q i = Q i ( p, q )

that is explicitly:
we say:
PR22 A canonical transformation exists iff there is a matrix X such that
t

X J 1 X = J 1

( or t X J X = J )

where
p

x P
X = B =

y q
P

Q
.
q

Proof. First, we deduce successively


From (i):

pi
Q j

2 S1
Q j q i

Pj
q i

from (ii):

pi
2 S2
Q j
=
=
,
Pj Pj q i
q i

from (iii):

Pj
2 S3
q i
=

=
,
Q j
Q j pi pi

(6-46)

424

from (iv):

Chapter 6

2 S4
q i
Q j
=
=
.
Pj
Pj pi
pi

x A
Second, we denote the matrix B by X BA .
y

( )

We have obviously:
x A y B
y x
B

= X BA ( X 1 ) CB = AC .

For instance,
p1 y B
p1 y B
= 1, ,
= 0,
y B p1
y B q 3

We emphasize that

(X )

1 A
B

y A
= B .
x

Now let us prove that


t

X 1 J 1 = J 1 X .

Indeed, we have:
t

X 1

p
=
P

p
Q

and
t

X 1 J 1

p
=
Q

P q

p P
=
P p

q P

Q
.
p

Since
q

P
1
J X =
p

Q
,
p

we have proved that


J 1 X = t X 1 J 1

and thus
t

X J 1 X = J 1 .

We note that such a fundamental condition of canonical transformation exists whatever the
generating function we choose.

Hamiltonian mechanics

425

In addition, there exists an essential structure of group, namely:


PR23 The set of matrices of canonical transformations on the phase space has a structure of
group.
Proof. (i) The composition of any two matrices of the set (matrix multiplication) is a matrix
of the set.
Indeed, given matrices X and Y such that
t

XJX =J,

( XY ) J ( XY )= t Y t X J X Y = t Y J Y = J .

Y JY = J ,

we have

(ii) The associativity results from the associative law of matrix multiplication.
(iii) The unit element is the unit matrix I.
It is such that
t
IJI=J.
(iv) Every element X of the set has an inverse, namely:
X 1 = J 1 t X J .

Indeed, we have:
X 1 X = J 1 t X J X = J 1 J = I

and X 1 is an element of the set because


t

( J 1 t X J ) J ( J 1 t X J ) = t J ( X t J 1 t X ) J = t J ( X J J X 1 J 1 ) J

= t J J 1 J = J .

4. HAMILTON-JACOBI EQUATION

We know that the solution of the canonical system of motion equations is obvious
when all the coordinates are cyclic. The analytical integration method of Jacobi is famous for
solving insoluble problems by the Lagrangian or Hamiltonian formalisms. This method based
on Hamilton works is very effective in celestial mechanics, in perturbation problems and
other areas.
4.1

HAMILTON-JACOBI EQUATION AND JACOBI THEOREM

The canonical transformations are particularly useful insofar as the transformed


Hamiltonian H is simpler than the original Hamiltonian H. A clever man as Jacobi thought to

426

Chapter 6

choose a zero transformed Hamiltonian in order that the solutions P i and Q i of new
canonical equations

H
Pi = i
Q

H
Q i =
Pi

be constants. Therefore there are 2n (independent) first integrals of motion of the


representative point in the phase space. The Hamilton-Jacobi equation is obtained from such a
judicious canonical transformation.
If the transformed Hamiltonian H is zero then the existence criterion of a canonical
transformation for a generating function of type S (t , q, Q ) is
pi dq i Pi dQ i = H dt + dS .

Identification implies the 2n equations


pi =

S
q i

Pi =

S
Q i

and the famous nonlinear Hamilton-Jacobi equation




S
S
+ H (t , i , q i ) = 0 .
t
q

(6-47)

In the Jacobian theory, the 2n equations of Hamilton being


Pi = 0

Q i = 0 ,

then the 2n (independent) first integrals of motion of the representative point in the phase
space are

Pi (t , pi , q i ) = ai

Qi (t , pi , q i ) = bi

where ai and b i are 2n arbitrary constants, the position of the index of bi doing not matter.
Given an orbit of a descriptive point in the phase space, the values of parameters ai and b i
are the motion constants.
In mechanics, only the search of a complete integral of the Hamilton-Jacobi partial
differential equation is interesting (and not the general solution).
D

A complete integral of the Hamilton-Jacobi equation is a solution which depends on


as many arbitrary independent integration constants as coordinates q i and t, on q i
and t.

However, the Hamilton-Jacobi equation shows S through partial derivatives. Therefore, one
of the arbitrary constants is necessarily additive (since it doesnt alter the partial
derivatives).This additive constant plays no role in the canonical transformation (only the
partial derivatives of S are present) and we express precisely:

Hamiltonian mechanics
D

427

A complete integral of the Hamilton-Jacobi equation is a solution of the equation


depending on as many nonadditive independent constants as coordinates q i , on q i
and t.

It is denoted by
S (t , q i , b i )

by taking the new coordinates Q i as arbitrary constants.


Now, the n relations
pi =

S (t , q i , b i )

q
at the initial instant are n equations linking the n constants b i and initial values of pi and q i
together. So, they allow to calculate the integration constants from particular initial
conditions.
i

Next, the n following equations


ai = Pi =

S (t , q i , b i )
i
b

provide the various ai from the initial conditions since the right-hand member is known from
initial values of q i .
Therefore, the problem is solved because the last equations let us obtain the coordinates q i
from initial conditions and time:
q i = q i (t , ai , b i ) .

It is the powerful integration method of Jacobi.


Remark 1. Dont forget the existence condition of every canonical transformation:

det (

2S
) 0.
q i b j

Remark 2. The notation of a generating function by S (t , q i , Q i ) is fully justified because we


have ( Q i being constants):
dS = pi dq i H dt = L dt .

Thus S (corresponding to a complete integral of the Hamilton-Jacobi equation) is nothing else


than the action integral S = L (t , q, q )dt , where this integration is along the solution q(t).
With each complete integral is associated a class of orbits of a representative point in the
phase space fitting the least action principle.

428

Chapter 6

Jacobis theorem

PR24 If

S (t , q i , b i ) is a complete integral of the Hamilton-Jacobi equation, then the

functions pi (t , ai , b i ) and q i (t , ai , b i ) obtained by solving the 2n equations


pi =

ai =

S (t , q i , b i ) ,

S (t , q i , b i )
i
b

constitute the general integral of Hamiltons canonical equations.


Proof. On the one hand, by considering pi =

S (t , q i , b i ) , the derivative of

S
S
+ H (t , q i , i ) = 0
t
q

with respect to parameters b i leads to


H 2 S
2S
= 0.
+
t bi p j q j bi

(1)

On the other hand, the total derivative of ai with respect to time, namely ai = 0 , is written:
2S
2S
+
q j = 0
bi t bi q j

(2)

(since bi = 0 ).

Subtracting (1) and (2), we obtain:


2S
H
(q j ) = 0
j
i
q b
p j
and the canonical transformation condition det (

2S
) 0 leads to n Hamiltons canonical
q j bi

equations:
q j =

H
.
p j

Now, on the one hand, the derivative of the Hamilton-Jacobi equation with respect to
coordinates q i leads to
2S
H H 2 S
+
+
= 0.
t q i q i p j q j q i
On the other hand, the total derivative of pi =
p i =

(3)

S
with respect to time is
q i

2S
2S
+
q j .
i
i
j
q t q q

(4)

429

Hamiltonian mechanics

Subtracting (4) and (3), we obtain:


p i =

H
2S
H
+
(q j )
i
i
j
q q q
p j

which implies the n other canonical equations:


p i =

H
.
q i

Remark. We have shown a canonical transformation (ai , b i ) ( pi , q i ) and the general


integral of the canonical equations of Hamilton:

pi = pi (t , ai , b i ) ,

q i = q i (t , ai , b i ) .

Any other general integral


pi = pi (t , ci , d i ) ,

q i = q i (t , ci , d i )

(5)

doesnt define a canonical transformation and doesnt allow obtaining a complete integral of
the Hamilton-Jacobi equation.
However, there is an exception, namely:
If the constants ci and d i are respectively the various initial values pi0 and q0i , then the
previous general integral shows a canonical transformation (ci = pi0 , d i = q0i ) ( pi , q i ) :
pi = pi (t , pi0 , q0i ) ,
Indeed, the general integral (5) with

q i = q i (t , pi0 , q0i ) .

D( pi , q i )
0 leads to the system
D(ci , d i )

ci = ci (t , pi , q i ) ,

d i = d i (t , pi , q i )

of 2n first integrals of Hamilton canonical equations and thus the following first
integrals:{ ci , c j }, { ci , d j } and { d i , d j }.
We immediately have:
(ci , d j )t 0 = i j ,

(ci , c j )t 0 = 0 ,

( d i , d j )t 0 = 0 ,

{ ci , d j } t 0 = i j ,

{ ci , c j } t 0 = 0 ,

{ d i , d j } t0 = 0

and since the Poisson brackets are first integrals we deduce (for every time):
{ ci , d j } = i j ,

{ ci , c j } = 0 ,

{ di,d j } = 0 .

Therefore, the general integral (5), where ci and d i are respectively the initial values of
various pi and q i , is a canonical transformation.

430

4.2

Chapter 6
SEPARABILITY

Obtaining a complete integral of the Hamilton-Jacobi equation is often a fastidious


even unsolved problem; nevertheless, searching this integral is sometimes made easy.
We restrain our study to the (frequent) case of conservative material systems with
scleronomic Hamiltonian.
In the case of a scleronomic Hamiltonian
i = 1,..., n

H ( pi , q i ) = E

the Hamilton-Jacobi equation


t S + H (

S i
,q ) = 0
q i

has a complete integral of type


S = E t + S (q 1 ,..., q n , E , b 2 ,..., b n )

(6-48)

where t S = 0 .
Therefore, the Hamilton-Jacobi equation becomes:
H(

S i
,q ) = E .
q i

(6-49)

The theorem of Jacobi immediately leads to the general solution of the canonical equations of
Hamilton:
pi =

S
q i

a1 =

S
S
=t
E
E

aj =

S
b j

i = 1,..., n ,
( or t t 0 =

S
),
E

j = 2,..., n .

Example. In the case of the simple harmonic oscillator, the Hamilton-Jacobi equation is
S
obtained by replacing p by
in the Hamiltonian, that is:
q
1 S 2 k 2
S
( ) + q =0
+
2
t 2m q

where q is the displacement and k the stiffness of the spring.


This Hamilton-Jacobi equation is easily solved by considering
S = S (q, b) b t
where b is the energy of the simple harmonic oscillator.

Hamiltonian Mechanics

431

The Hamilton-Jacobi equation


H(

S
, q) = b
q

is here
1 S 2 k 2
( ) + q = b.
2m q
2
Thus the complete integral is

S = km

2b

q 2 dq b t .

We have:
a=

S
km
=
b
k

ta =

q=

dq
2b k q 2

+t

m
k
cos 1 (
q)
k
2b

2b
k
cos(
(t a))
k
m

The constants a and b are linked to initial conditions.


At initial instant (t = 0) the oscillating point of mass m is relaxed at equilibrium position q0
with p0 = 0 .
Given the initial conditions, the Hamilton-Jacobi equation

S 2
) = 2m b km q 2
q

implies
0 = p0 = (

S
)0 = 2m
q

k
2

q02

and we find again an obviousness: the parameter b is the (initial) mechanical energy
b = k q02 = m 2 q02 = E .
2

Therefore, we have:
q = q0 cos( (t a ))
and the initial condition q(0) = q0 implies the constant a must vanish.
In conclusion, S is the generating function of a canonical transformation and is expressed by
S = m
(given the initial conditions).

q02 q 2 dq m
2 q02 t
2

432

Chapter 6

PR25 A scleronomic mechanical system with a cyclic generalized coordinate, e.g. q1 , shows
a complete integral of the Hamilton-Jacobi equation of type
S = E t + a1 q 1 + S (q 2 ,..., q n ; a1 , E , b 2 ,..., b n 1 )

(6-50)

where a1 is an arbitrary constant.

Proof. The scleronomic character explains the first term. In addition, if q1 is a cyclic
coordinate we have:
L
=0
q1

H
= 0 = p1 .
q1

S
conjugate to q1 is an arbitrary constant a1 .
1
q
The complete integral is thus of type (6-50). Since it must verify the Hamilton-Jacobi
equation, then the function S must satisfy:

The generalized momentum p1 =

H ( q 2 ,..., q n , a1 ,

S
S
,..., n ) = E .
2
q
q

In the complete integral (6-50) the variables t and q1 are separated from the other variables.
The reader will generalize if t , q1 ,..., q k are cyclic coordinates, with H (q k +1 ,..., q n , p1 ,..., pn ) .
D

A scleronomic conservative mechanical system is called separable


if the
corresponding Hamilton-Jacobi equation shows a complete integral S of type:

S = E t +

S i ( q i ; b)

(6-51a)

i =1

where b denotes the set of arbitrary constants.


This complete integral will be also denoted simply by:
n

S = E t + Si ( q i ) .

(6-51b)

i =1

where only the corresponding variable qi which appears in each Si is indicated.


Make clear that the separability existence is dependent on the choice of generalized
coordinates.
Remark. Solving the Hamilton-Jacobi partial differential equation is not a priori more simple
than solving Hamilton canonical equations, but separability leads to quadratures!
Conditions for separability

There are cases where a Hamiltonian system may be quickly solved by separation of
variables; they are the most simple Hamiltonian cases.

433

Hamiltonian Mechanics

Stckel1 showed a general form of the Hamiltonian for which we can have separation of
variables. Historically, conditions for the separability of variables were obtained by Morera
and DallAqua, and works of Levi-Civita were essential. Later, Weinacht2 found all the
coordinate systems of two and three dimensions for which separation is possible. At present,
the study of separable systems has numerous applications, as for instance in astronomy.
A necessary and sufficient condition for the separability can be shown.
Let
n

S = Et + S i (q i )
i =1

be a complete integral of the Hamilton-Jacobi equation. It is such that

H (q i ,

(where the notation

dS i

dS i
dq i

)=E

i = 1,..., n

emphasizes the derivation with respect to the only concerned

dq i
generalized coordinate),
and thus, k {1,..., n} :
dH
dq k

H
q k

H d 2 S k
=0
p k (dq k ) 2

(no summation).

d 2Sk
k 2

(dq )

H
q

H
p k

H
o.
p k

by assuming k :

For every r k , we have:


d
dq

H
q

H
)=0
p k

H
H dp r H H H
H dp r
( k )+
( k) r]
k[ r(
)+
(
)
]=0
r
p r q dq p k q q p k
p r p k dq r
q q

and since
dp r
dq

H
q r

H
,
p r

we obtain the conditions for separability:

(
1

2 H H
2 H H H
2 H H
2 H H H

)
(
)
= 0.
q r q k p r p r q k q r p k
q r p k p r p r p k q r q k

Stckel, P. 1890, Math. Ann. 35, 91.

1893, ibid. 42, 537.


2
Weinacht, J. 1923, Math. Ann. 90, 279.

(6-52)

434

Chapter 6

We consider Hamiltonians H = T + V such that


T = 12 a ij (q) pi p j ,

V = V (q ) .

n(n 1)
necessary and sufficient conditions for the separability (6-52) are
2
polynomials of fourth degree in p, where the coefficients of the terms of fourth, second and
zero-th degrees are equal to zero:

So, the

2T T T
2T T T
2T T T
2T T T

+
= 0,
q r q k p r p k q r q k p r p k p r q k q r p k p r p k q r q k

(1)

2T T V
2V T T
2T T V

q r q k pr pk pr q k pk q r q r pk pr q k

(2)

T V
2T T V
( r k + k r) =0,
pr pk q q
q q

2T V V
V V
= a kr k
=0
k
r
p r p k q q
q q r

(3)

( r k ).

Equation (1) determines the form of T , equations (2) and (3) determine V.

Stckels case

The important case where the various akr vanish for k r (the Hamiltonian containing
a sum of n momenta squared) was studied by Stckel.
We say:
PR26 A material system such that
H=

1
2

f i (q) pi2 + V (q)


i =1

is separable iff there are:

( )

(i) a nonsingular matrix uij of which each element depends on the only variable qi
and such that
n

f i uij = 1j ,
i =1

(ii) functions vi of the corresponding variable qi such that


n

f i vi = V

(potential).

i =1

Proof. First, from the separability, we are going to deduce (i) and (ii).

435

Hamiltonian Mechanics
n

A complete integral S = S i (q i ; b) verifies the Hamilton-Jacobi equation H = E ( = b1 ) ,


i =1

that is:
n

1
2

dS

f i ( dq ii ) 2 = b1 V .
i =1

Derivations with respect to successive parameters bi lead to

fi

dS i dS i
(
) =1 ,
dq i b1 dq i

fi

dS i dS i
(
)=0
dq i b j dq i

i =1
n

i =1

j = 2,..., n .

We note that each fi depends on the only corresponding variable qi and the system of the
previous equations can be written:
n

i =1

f i u ij = 1j

where each u ij depends on the only corresponding generalized coordinate qi.


In addition, the expression
n

V = b1 12 f i (
i =1

dq

) 2 = 1j b j 12 f i (
i

= f i (uij b j 12 (
i =1

is really of type

dS i

i =1

dS i
dq

dS i
dq

)2

)2 )

f i vi where each vi only depends on the corresponding variable qi.

i =1

Second, if by hypothesis we have:


n

i =1

f i u ij = 1j ,

f i vi = V

i =1

then the Hamilton-Jacobi equation H = b1 , that is


n

1
2

i =1

fi (

S 2
) + V = b1 ,
i
q

is written
n

i =1

f i ( 12 (

S 2
) + vi u ij b j ) = 0 .
i
q

436

Chapter 6

Therefore, the Hamilton-Jacobi equation has solutions of the following type:

S = S1 (q1 ; b) + ... + S n (q n ; b)
by choosing, for each i:

S 2
) = 2 (u ij b j vi )
i
q
= i (q i ; b).

A complete integral is thus written:


n

S = E t +

i =1

i dq i .

We note that the Jacobis theorem leads to the general solution of Hamiltons canonical
equations, that is
pi = i

(1)

and
n
u1
a1 = S = t 22 i dqi
E
i =1
i

also written:
n

t t0 =

i =1

ui1

dq i ,

(2)

and
n

a j =

i =1

u ij

dq i

j = 2,..., n,

(3)

these last equations (3) determining the trajectories.

Example. We consider a system with two degrees of freedom of which the Hamiltonian has
the form
H = 12 [ A( x, y ) p x2 + B( x, y ) p 2y ] + V ( x, y )

where x, y are generalized coordinates and px, py the conjugate momenta.

(i) If the system is separable, characterize the functions A, B, V and find the expression of a
complete integral.
(ii) Deduce the equations of motion.
Answer. (i) A complete integral of the separable system is of type
S = S1 ( x, E , b) + S 2 ( y, E , b)

and thus
1
2

[ A( x, y ) (

dS
dS1 2
) + B ( x, y ) ( 2 ) 2 ] = E V .
dx
dy

437

Hamiltonian Mechanics

By letting
f ( x, E , b) = 12 (

dS1 2
) ,
dx

g ( y, E , b) = 12 (

dS 2 2
) ,
dy

we obtain
A( x, y ) f ( x, E , b) + B( x, y ) g ( y, E , b) = E V .

(1)

So, we have the system of equations


A

f
g
+B
=1 ,
E
E

f
g
+B
= 0.
b
b

Since S is a complete integral, we have:


2S
xE
D= 2
S
yE

2S
xb
2S
yb

and thus
dS1 dS1
g
(
)
dx E dx
E
=
dS1 dS1
g
(
)
b
dx b dx

f
E
=
f
b
=

dS 2 dS 2
(
)
dy E dy
dS 2 dS 2
(
)
dy b dy

dS1 dS 2
D 0.
dx dy

So, the previous system of equations has one solution


1
0
A =

B =

f
E
f
b

g
E
g
b

1
,
f f E g E

(
)
b f b g b

1
0

1
.
g f E g E
(
)

b f b g b

By letting
P=

1
,
f b

Q=

1
,
g b

X =

f E
,
f b

Y =

g E
g b

438

Chapter 6

where P and X do not depend on y and Q and Y do not depend on x, we have so obtained:
A=

P
,
X +Y

B=

Q
.
X +Y

Equation (1) is thus written:


V = E Af Bg =

EX + EY Pf Qg
.
X +Y

By letting the function (independent on y)

= EX Pf ,
and the function (independent on x)

= EY Qg ,
we have:
V=

+
X +Y

and
H=

1
+
( P p x2 + Q p 2y ) +
.
X +Y
2( X + Y )

Conversely, given such a Hamiltonian there is separability, the corresponding HamiltonJacobi equation is
1
P S
Q S
[ ( )2 + ( )2 + + ]= E .
2 y
X + Y 2 x
A complete integral of type S = S1 ( x, E , b) + S 2 ( y, E , b) is such that S1 is obtained from
P dS1 2
(
) = EX + b
2 dx

and S2 is obtained from


Q dS 2 2
(
) = EY b ;
2 dy

that is:
2
( EX + b dx +
P

S=

(ii)

We have successively:

px =

S
=
x

2
( EX + b)
P

py =

S
=
y

2
( EY b)
Q

2
( EY b) dy .
Q

439

Hamiltonian Mechanics

t t0 =
a=

X dx
Y dy
S
,
=
+
E
2 P( EX + b)
2Q( EY b)

S
dx
=
+
b
2 P( EX + b)

dy
2Q( EY b)

From the two previous results, we deduce:

dx
2 P( EX + b)

dy

2Q( EY b)

and

dt =

dx

2 P( EX + b)

dy

2Q( EY b)

( X + Y ) dx
2 P( EX + b)

So, the trajectory and the motion are determined by the following equations:

dx
2 P( EX + b)

dy
2Q( EY b)

dt
.
X +Y

5. EXERCISES
Exercise 1.

Write down the Hamiltonian for a free particle in cylindrical coordinates. Write the
canonical equations of Hamilton and the Liouville-Boltzmann equation. Show that this last
equation leads to Hamiltons equations.

Answer. By considering the generalized coordinates


q1 = r ,

q2 = ,

q3 = z ,

the Lagrangian for the free particle is

L = T V = 12 (r 2 + r 2 2 + z 2 ) V (t , r , , z ) .
Since the generalized momenta are
pr =

L
= r ,
r

the Hamiltonian per unit mass is written

p =

L
= r 2 ,


pz =

L
= z ,
z

440

Chapter 6

H = pi q i L
= 12 ( p r2 +

p2
r2

+ p z2 ) + V (t , r , , z )

The canonical equations of motion are

 =

r = p r ,
p r =

H p2 V
,
= 3
r
r
r

p
r2

p =

z = p z ,

V
,

p z =

V
.
z

The Liouville-Boltzmann equation is written:


f p f
f p2 V
f V f V
df
f f
=
+
pr +
+
pz +
( 3
)

= 0.
2
r
z
pr r
r
p p z z
dt t r

This linear homogeneous equation in f leads to the system of canonical equations:


dp
dp
dp z
dt dr d
dz
=
=
=
= 2 r
=
=
.
p
1
pr
pz
p V V V

z
r2
r 3 r
Exercise 2.
From the well-known equations of motion of a particle (of a system) written in
generalized coordinates:
qi + ijk q j q k = a i
where a i = g ik

V
,
q k

(i) express the motion equations in cylindrical coordinates and the Liouville-Boltzmann
equation,
(ii) express the motion equations in spherical coordinates and the Liouville-Boltzmann
equation,
(iii) find again the corresponding expressions of Lagrangian, Hamiltonian and canonical
equations in spherical coordinates.
Answer. (i) Given
ds 2 = dr 2 + r 2 d 2 + dz 2 ,

the only nonzero Christoffel symbols are


1
.
r

= r ,

r = r =

q1 = r = R ,

q 2 =  =
,
r

We denote
q 3 = z = Z ,

where R is the velocity component in the r-direction, is the velocity component


perpendicular to the (r , z ) plane and Z is the velocity component along the z-axis.

441

Hamiltonian Mechanics
The motion equations are the following:
1
q1 + 22
q 2 q 2 = g 11

V
q 1

2 V

R=

,
r
r

and
2 1 2
2
q2 + 12
q q + 21
q 2 q 1 = g 22

1 V
d 2
( ) + r = 2
dt r
r r
r

 = R 1 V ,

r
r

V
q 2

and
q3 = g 33

V
q 3

V
Z =
.
z

With the six variables r , , z , R, , Z , which are not canonically conjugate, the equations of
the motion of the particle are simplified.
For a phase density f (t , r , , z , R, , Z ) , the Liouville-Boltzmann equation is
f
f f
f
2 V f
R 1 V f V f
+R +
+Z
+(

(
+

)
)
= 0.
t
r r
z
r R
r
r
r z Z

(ii) Given
ds 2 = dr 2 + r 2 d 2 + r 2 sin 2 d 2 ,

the only nonzero Christoffel symbols are


r

= r sin 2 ,

= r ,

r = r =

r = r =

r
r


= sin cos ,

=
= cot ,

We denote
q 1 = r = R ,

q 2 =  =
,
r

q 3 =  =

r sin

442

Chapter 6

where R is the velocity component in the direction of the radius vector simply called the radial
velocity, is the velocity component along the meridian and is the velocity component
along the circle of colatitude.
The motion equations are the following:
1
q 3 q 3 = g 11
q1 + 122 q 2 q 2 + 33

V
q 1

2 + 2 V
R =
,

r
r

and
2 1 2
2
2
q 3 q 3 = g 22
q2 + 12
q q + 21
q 2 q 1 + 33

V
q 2

 = R + cot 1 V ,

r
r
r
2

and
3 1 3
3
3
3
q3 + 13
q q + 31
q 3 q 1 + 23
q 2 q 3 + 32
q 3 q 2 = g 33

V
q 3

 = R cot 1 V .

r
r
r sin

With the six variables r , , , R, , , which are not canonically conjugate, the equations of
the motion of the particle are simplified.
Given the phase density f (t , r , , , R, , ) , the Liouville-Boltzmann equation is thus
written:

f
f f
f
2 + 2 V f
+(

)
+R +
+
t
r r r sin
r
r R
R 2
1 V f
R
1 V f
+ (
+
cot
)
+ (

cot
)
= 0.
r
r
r
r
r
r sin
(iii)

Given a mass m, we have immediately:


L=

m 2
(r + r 2 2 + r 2 sin 2  2 ) V (t , r , , ) ,
2

H = pi q =
i

1
2m

( p r2

p2
r2

In generalized coordinates, we note that


L=m
g q i q j V (t , q ) ,
2 ij

H = 21m g ij pi p j + V (t , q) .

p2
r 2 sin 2

) + V (t , r , , ).

443

Hamiltonian Mechanics

The canonical equations

p i = { pi , H }

q i = q i , H ,

are written as it follows:

r =

p
H
= r ,
p r
m

 =

p
mr 2

 =

p
mr 2 sin 2

p2
p2
H
V
=
+

p r =
,
3
3
2
r mr
mr sin r
p =

p2
mr sin

p =

cot

V
,

V
.

We note that the generalized momenta are written with the variables R, , as follows:

p r = mr = mR ,
p = mr 2 = mr ,
p = mr 2 sin 2  = mr sin .

Exercise 3.
Is the diffeomorphism
P = q cot p ,

Q = ln(q 1 sin p)

a canonical transformation on the 2-dimensional phase space ?


Answer. Yes it is, because
cos p
q
dp q 1 dq )
dp + cot p dq ) (
2
sin p
sin p
1
= ( 2 cot 2 p ) dp dq
sin p

dP dQ = (

= dp dq .

Exercise 4.
Find the motion equation of a simple harmonic oscillator by introducing the generating
function
S1 = 12 m (q ) 2 cot Q ,

where q is the displacement and 2 = k m .

444

Chapter 6

Answer. We have:
p=

S1
= m q cot Q ,
q

P=

S1 1 (q) 2
= m
.
Q 2 sin 2 Q

This last relation leads to


2P
sin Q
m

q=
and thus

p = 2m P cos Q .

The Hamiltonian is invariable under the transformation because the time doesnt explicitly
appear in the generating function.
Since
p=

T
= m q
q

and
V =

k
2

(q) 2 ,

the Hamiltonian is written:


H = T +V =

( p) 2 k
+ (q) 2 = P cos 2 Q + P sin 2 Q = P .
2m
2

Since the coordinate Q is cyclic, we conclude that the conjugate momentum P is constant, it
is E .
The motion equation is
H
=
Q =
P

which implies
Q = t + c ,

where c is an integration constant (fixed by the initial conditions).


We find the well-known displacement
q=

2E
m 2

sin ( t + c) .

445

Hamiltonian Mechanics
Exercise 5.

In the central force problem, find a complete integral of the Hamilton-Jacobi equation
and the general solution of motion equations.
Answer. The Hamiltonian is well-known:

H = 21m ( pr2 + p2 r 2 ) + kr .
The Hamilton-Jacobi equation is

k
S 2 1 S 2
t S + 1 [( ) + 2 ( ) ] + = 0 .
2 m r
r
r
We search a complete integral of type
S = E t + S1 (r ) + S 2 ( )

which implies
1 [( S ) 2 + 1 ( S ) 2 ] + k = E
2m r
r 2
r

that is

dS 2 2
k
dS 2
) = r 2 [ 2m E 2 m ( 1 ) ] .
d
r
dr

Since the two members are functions of respective and r, we obtain:


dS 2
= A.
d

This expected result, S2 = A , implies:


r 2 [ 2m E 2 m

k
dS 2
( 1 ) ] = A2
r
dr

S = E t + A

2mE 2m

k
r

A2
dr .
r2

The general solution is such that


b1 =

S
= t m
E

b2 =

S
=
A

dr
2mE 2m k r A2 r 2

A dr
2

2mE 2m k r A2 r 2

thus, for a trajectory through (r0 , 0 ) at instant t0 , the previous equations are well-known:
t t0 =

m
2

r0

dr
E k r A2 2mr 2

446

Chapter 6
r

dr

r0

r 2 2mE 2m k r A2 r 2

0 = A

These equations give the positions ( by (r ) ) and position instants.

Exercise 6.

A particle moves in a vertical plane under the action of its weight mg without friction.
This plane is constrained to rotate about a vertical axis with constant angular velocity .
(i)
Calculate the Hamiltonian.
(ii)
Find the Hamilton-Jacobi equation and a complete integral.
(iii) Determine the general solution of Hamilton canonical equations.
Answer.
(i) By introducing the cylindrical coordinates r , , z and knowing that  = , the Lagrangian
is written:
L=

m 2
m
(r + r 2  2 + z 2 ) mg z = (r 2 + z 2 + r 2 2 ) mg z .
2
2

There are two degrees of freedom, the generalized coordinates are r and z.
Since the generalized momenta are
pr =

L
= m r ,
r

pz =

L
= m z ,
z

the Hamiltonian is immediately:


H = pi q i L =

1
m
( pr2 + p z2 ) r 2 2 + mg z .
2m
2

(ii) The Hamilton-Jacobi equation is


S
S 2 S 2
1
+
( ( ) + ( ) ) m r 2 2 + mg z = 0 .
2
t 2m r
z

By separating the variables t , r , z :


S = E t + S1 (r ) + S2 ( z ) ,

the Hamilton-Jacobi equation becomes:


dS 2 dS 2
1
( ( 1 ) + ( 2 ) ) m r 2 2 + mg z = E
2m dr
2
dz

or

1 dS1 2 m 2 2
1 dS 2
(
) r ) + ( ( 2 ) + mgz ) = E .
2m dr
2m dz
2

So, the first term only depends on variable r and the second term only on variable z; they are
thus constants.

447

Hamiltonian Mechanics

By putting
1 dS1 2 m 2 2
(
) r = E1 ,
2m dr
2
1 dS 2 2
(
) + mg z = E2
2m dz

with E1 + E2 = E ,
we have a complete integral immediately:
S = ( E1 + E2 ) t 2mE1 + m 2 r 2 2 dr 2mE2 2m 2 g z dz .
(iii) The general solution of motion equations follows from
a1 =

S
,
E1

a2 =

S
E2

that is
a1 = t

m dr
2mE1 + m r
2 2

a2 = t

m dz
2mE2 2m 2 g z

Exercise 7.

The Hamiltonian of a system with two degrees of freedom is


H=

1
( 12 ( g p12 + f p 22 ) + g + f )
f +g

where f and are functions of the generalized coordinate q1, and g and are functions of the
coordinate q2.
Find a complete integral of the Hamilton-Jacobi equation.
Answer. It is a Stckels case of separation of variables and we have:
S = Et + S1 (q 1 ) + S 2 (q 2 ) .

The Hamilton-Jacobi equation is written:


dS
dS
1 1
[ 2 ( g ( 11 ) 2 + f ( 22 ) 2 ) + g + f ] = E
f +g
dq
dq

that is
1 ( 1 ( dS1 ) 2 + 1 ( dS 2 ) 2 ) + + E E = 0 .
2 f dq1
g dq 2
f g g f

By introducing the constant C, the separation is expressed as


1 dS1 2 E
(
) + =C,
f
f
2 f dq1
1 dS 2 2 E
(
) + = C .
2 g dq 2
g g

448

Chapter 6

So, we have
S1 = 2Cf 2 + 2 E dq1 ,
S 2 = 2Cg 2 + 2 E dq 2

and thus

S = Et + S1 + S 2 .

Exercise 8.
Find a complete integral of the Hamilton-Jacobi equation of a spherical pendulum of
length R. Establish the general solution of canonical equations of Hamilton.
Answer. In this problem of two degrees of freedom the Lagrangian is written
L=

m
2

R 2 ( 2 +  2 sin 2 ) + mg R cos

where the two generalized coordinates are the colatitude and the longitude ,and the
Hamiltonian is
p2
1
2
H ( p , p , ) =
(
p
+
) mg R cos .

2
2mR
sin 2
The Hamiltonian-Jacobi equation is written:
t S +

1
2 mR

((

1
S 2
S 2
( ) ) mgR cos = 0 .
) +
2
sin

In this scleronomic problem where is a cyclic coordinate, we search a complete integral of


type:
S = E t + c + ( ) .

The Hamilton-Jacobi equation that is


1
2mR

((
2

d 2
c2
) +
) mgR cos = E
d
sin 2

lets us obtain .
Thus a complete integral is
S = E t + c 2mR 2 ( E + mgR cos ) c 2 sin 2 d .
The general solution of motion equations follows from
m R 2 d

a1 =

S
=t
E

a2 =

S
= c
c
sin 2

2mR 2 ( E + mgR co ) c 2 sin 2

= t t0

d
2mR ( E + mgR cos ) c 2 sin 2
2

449

Hamiltonian Mechanics

Remark 1. The first part of the general solution shows the horary law, the second leads to
( , E , c, a2 ) .
Remark 2. We can immediately verify that:
2S
E

S
)=
ai q j
2S
c
2

det(

2S
E

d
E d
0.
d
c d

0
=

2S
c

Exercise 9.

A particle of mass m and position vector ox moves in a vertical plane under the action
of its weight mg and a central force inversely proportional to the square of the distance from
the center r = ox .
(i) Express the Hamiltonian in function of variables
X =r+z ,

Y =rz

where z is the height of the particle relative to o.


(ii) Find the general solution of the canonical equations of Hamilton.
Answer. (i) From the expression of the potential
mk 2
V = mgz
r
we deduce
L = T V =

m
2

kR

( x 2 + z 2 ) mgz + m

k2
r

and thus
H = p x x + p z z

x = r2 z2 =

XY

( x 2 + z 2 ) + mgz

mk 2
x +z
2

Since

we deduce
L=

z=

X Y
2

m ( XY + XY ) 2
X Y 2
( X Y ) 2mk 2
(
) ) mg
+(
+
X +Y
2
4 XY
2
2
mg

( X 2 Y 2 ) 2mk 2
m
X 2 Y 2
= ( X + Y )(
+
) 2
2
4 X 4Y
X +Y
The generalized momenta
pX =

L
,
X

pY =

L
Y

450

Chapter 6

imply
4X pX
,
X =
m( X + Y )

Y =

4Y pY
.
m( X + Y )

Therefore, the Hamiltonian is written:


H = p X X + pY Y L
mg

2
m( X + Y )

( X p X2 + Y pY2 ) + 2

( X 2 Y 2 ) 2mk 2
X +Y

In this scleronomic case, where X and Y are separable variables, the complete integral is of
type
S = Et + S1 ( X ) + S 2 (Y )
and thus

mg
( X 2 Y 2 ) 2mk 2
dS 2 2
dS1 2
(X ( ) + Y ( ) ) + 2
=E
m( X + Y )
dY
X +Y
dX
2

that is:
dS
dS
2
2
mg 2
mg 2
X EX + Y ( 2 ) 2
Y EY = 2mk 2 .
X ( 1 )2 +
m
dX
2
m
dY
2

(ii) By letting
2
m

X (

dS 1 2
mg
) +
X
dX
2

EX = C

and
2
m

Y(

dS 2 2 mg 2
)
Y EY = 2mk 2 C ,
dY
2

we obtain:
m C m 2 gX
+

dX = (1)
2
2 X
4

S1 =

mE

S2 =

mE

and
2

m 2 k 2 mC m 2 gY

+
dY = (2) .
2Y
4
Y

Therefore, we obtain:
S = Et + (1) + (2)

and the general solution is given by


a1 =

S
,
E

a2 =

S
.
C

451

Hamiltonian Mechanics

Exercise 10.

In spherical coordinates r , , , let


2

p
p
H = 1 ( pr2 + 2 + 2 2
2m
r
r sin

e2 b2 2 2
eb
r sin
)+
p + eV (r )
2
8mc
2mc

be the Hamiltonian of a particle of charge e subject to a central electric field [ potential V (r ) ]


and to a constant magnetic field B = b 1z (along pole axis), c being the velocity of light.
(i) Deduce there is no separable solution of the Hamilton-Jacobi equation.
(ii) If the term (eb c) 2 is neglected, show the Hamilton-Jacobi is separable. Find a complete
integral and the general solution of Hamilton canonical equations.
Answer. (i) The coordinate being cyclic, let us try a separable solution of type
S = E t + R (r ) + ( ) + A

where the constant A is the generalized momentum associated with .


The Hamilton-Jacobi equation is written:
eb
dR 2 1 d 2
A
eb 2 2
1
(( ) + 2 ( ) + 2 2 ) +
A + eV (r ) = E
r sin
2
2m dr
r d
r sin
8mc
2mc
2

2 2

or
dR
d
A2
eb
e 2b 2 4 2
1
(r 2 ( ) 2 + ( ) 2 + 2 )
r sin = 0 .
A r 2 + e r 2 V (r ) E r 2 +
2m
dr
d
sin
2mc
8mc 2
Two successive partial derivatives lead to the following absurd result:
2
c 2b 2 4 2
(
r sin ) = 0 .
r 8mc 2
There is thus no separable solution.
(ii) If (

eb 2
) = 0 , then the Hamilton-Jacobi equation is written:
c
r 2 dR 2 eb
d 2
A2
1
2
2
2
( ( )
(( ) + 2 ) = 0 ,
A r + e r V (r ) E r ) +
2m dr
2mc
2m d
sin

where the first term depends only on r and the second term only on ; there are two (constant)
opposite terms:
r 2 dR 2 eb
( )
A r 2 + e r 2 V (r ) E r 2 = B ,
2m dr
2mc
A
1 d 2
(( ) +
) = B .
2m d
sin 2
2

Therefore, a complete integral is


S = E t

2mB
eb
A 2meV (r ) + 2mE + 2 dr
r
c

2mB

A2
d + A .
sin 2

452

Chapter 6

The general solution of motion equations is


a1 =

S
= t
E

a2 =

S
=
A

a3 =

m dr
eb
2 mB
A 2 meV ( r ) + 2 + 2mE
c
r

e b dr
2c

S
=
B

eb

2mB
A 2me V (r ) + 2 + 2mE
c
r

m dr
r

eb
c

A 2 meV ( r ) +

2mB

r2

+ 2mE

A sin 2 d

2mB A2 sin 2

d
2mB A2 sin 2

+.

BIBLIOGRAPHY

The books in question are essentially at the root of the theoretical mechanics teachings of the author for
undergraduate and third year engineering students.

Arnold, V., 1978, Mathematical methods of classical mechanics, MIR (Moscow, 1974), Springer
Graduate Texts in Math. N60 Springer-Verlag, New York.
Brousse, P., 1981, Mcanique analytique, Vuibert, Paris.
Chevalier, L., 1996, Mcanique des systmes et des milieux dformables, Ellipses-Marketing, Paris.
Flanders, H., 1963, Differential forms with applications to the physical sciences, Academic Press.
Goldstein, H., 1956, Classical mechanics, Addison-Wesley, Reading, Mass.
Kovalevsky, J., 1963, Introduction la mcanique cleste, Armand Colin, Paris.
Landau, L. and Lifshitz, E., 1960, Mechanics, Addison-Wesley, Reading, Mass.
Lichnerowicz, A., 1964, Elments de calcul tensoriel, A. Colin, Paris.
Mantion, M., 1981, Problmes de mcanique analytique, Vuibert, Paris.
Meriam, J., 1980, Dynamics, SI version, John Wiley & Sons.
Meriam, J., 1975, Statics, SI version, John Wiley & Sons.
Pars, L., 1965, A treatise on analytical dynamics, Heinemann, London.
Poincar, H., 1957, Mthodes nouvelles de la mcanique cleste, 3 vol.,Dover Publications.
Scheck, F., 1994, Mechanics from Newtons laws to deterministic chaos, Springer-Verlag.
Simon, K., 1979, Mechanics, Addison-Wesley, Reading, Mass.
Talpaert, Y., 1982, Mcanique analytique, vol. 2, Talpaert (Ed.).
Talpaert, Y., 1987, Mcanique gnrale et analytique, Ellipses-Marketing, Paris.
Talpaert, Y., 2000, Differential geometry with applications to mechanics and physics, Dekker,
New York.
Whittaker, E. T., 1970, A treatise on the analytical dynamics of particles and rigid bodies,
Cambridge University Press.
Wintner, A., 1947, The analytical foundations of celestial mechanics, Princeton University Press.

453

INDEX

absolute acceleration, 28, 304


absolute derivative, 215

calculus of variations, 212, 360


canonical equations, 395, 401, 412
canonical isomorphism, 166
canonical system, 396
canonical transformation, 416, 419-420, 423425
canonically conjugate momentum, 399
center of reduction, 14
central axis, 15, 41
central axis of inertia, 288
central ellipsoid of inertia, 288
central force field, 403, 405
change of basis, 140, 191, 312
change of cobasis, 140
change of natural basis, 198
characteristic equation, 282
characteristic function, 395
Christoffel formulae, 206
Christoffel symbols, 204
Christoffel symbols of first kind, 205
class of inertial frames, 306
cloud of possible states, 406
compatible with constraint(s), 81, 88, 90-91
complementary equation, 359, 402
complete integral, 426, 427
complete variation, 363
completely antisymmetric p-linear form, 178
completely symmetric tensor, 154, 155
component, 3
components of dynam, 5
compression, 48
concurrent (spatial forces), 44
cone of kinetic friction, 54
cone of static friction, 54
configuration space, 72, 74, 354
configuration spacetime, 358
conjugate tensor, 168
conjugate variables, 395
conservation of angular momentum, 318
conservation of mass (theorem), 311
conservative force field, 108
consistent with constraints, 81, 88, 90-91
constraint force, 34
constraint (particle subject to ), 86

( ) components, 208
absolute differential of ( ) components, 208

absolute differential of

1
0
0
1

absolute differential of a vector field, 207, 214


absolute frame of reference, 303-304
absolute velocity, 22
acceleration, 215
action and reaction (principle of), 39
action integral, 368, 378, 427
addition in L( E , F ) , 136
addition of dynams, 8
addition of p-forms, 185
addition of tensors, 156
adjoint Lagrangian, 359, 402
adjoint of a p-form, 219
adjoint of a q-vector, 218
admissible diffeomorphism, 376
affine space, 2
algebra, 413
alternation mapping, 178
angle of kinetic friction, 54
angle of static friction, 54
angular momentum, 307-308, 319-322
angular momentum theorem, 316
angular velocity, 18
angular velocity tensor, 312-314
angular velocity vector, 314
annular-linear constraint, 58, 67
antisymmetric tensor, 154
antisymmetrization, 178
axis of sliding dynam, 14

B
ball-and-socket joint, 59, 67
Bernoulli-Lagrange vector, 81
bilateral constraint, 75
body force, 49
brachistochrone problem, 367
bras, 137

455

456

constraint (system subject to a ), 75


constraints (system subject to ), 77
contact dynam, 48
continuous vector field, 92
contracted multiplication, 159
contracted product, 159, 172
contraction, 158
contravariant components, 171
contravariant-contravariant representation, 173
contravariant-covariant representation, 173
contravariant vector, 146
coordinate basis, 195
coordinate line, 193
coordinate system, 193
coordinates, 3, 192
coplanar forces, 43-44
Coriolis acceleration, 23, 39
Coriolis force, 323
Coulomb friction, 50
couple, 12, 13, 14, 16
covariant component, 170
covariant-contravariant representation, 173
covariant-covariant representation, 174

( ) -covariant derivative, 208


( ) -covariant derivative, 209
1
1

0
2

covariant derivative of t,, 211


covariant representation, 170
covariant representation of vector differential,
203
covariant vector, 146
covector, 136, 144, 219
curl of covector field, 228-229
curl of second order tensors, 229
curvilinear coordinates, 194, 197-199, 202
cyclic coordinate, 360, 404

D
dAlemberts principle, 342
dAlembert-Lagrange principle, 342
decomposable p-form, 183
density, 264
determinant, 217
diffeomorphism, 374
differential of function, 220
differential of vector, 195
dimension, 2
directional derivative of function, 221
directional derivative of tensor field, 224
divergence of tensor field, 227
divergence of vector field, 225

Index
domain of a material system, 263
double contraction, 160
double pendulum, 73
dry friction, 50
dual basis, 137
dual space, 137
dummy index, 139
dynam, 3-5
dynam of forces, 36, 95, 317
dynam of internal forces, 39
dynam of mechanical actions, 35
dynam of velocities, 21
dynam of virtual displacements, 93
dynam of virtual velocities, 93
dynamic dynam, 318

E
eigenvalue, 281
eigenvector of tensor, 281
Einstein summation convention, 318
elements of reduction, 5, 35
embedding, 67
equal tensors, 156
equality of dynams, 6
equilibrium, 38
equilibrium conditions, 40-42
equilibrium (particle in ), 98, 107, 109
equilibrium (rigid body in ), 100
equilibrium (system of particles in ), 107, 109
equilibrium (system of rigid bodies in ), 100
equiprojective field, 10, 20
equivalent systems of vectors, 4, 40
Euclidean space, 3
Euclidean vector space, 174
Euler angles, 82, 85
Eulers equation 212, 364, 365
Euler-Noether theorem, 374
Euler pendulum, 23
exterior algebra, 188
exterior calculus, 178
exterior form of degree p, 178
exterior multiplication, 186
exterior product of 1-form, 180
exterior product of p-form, 184, 186
exterior product of q-vectors, 188
exterior product space, 185
external constraint, 37
external forces of constraint, 37
external mechanical action, 34
extremal 365, 369-371
extremum principles, 360

457

Index
F
Fermats principle, 360
field of moments of dynam, 91
field of virtual velocities, 85
first integral, 310, 359, 366, 396, 413, 415
first variation, 364
flat mapping, 166
force, 300
force associated with q i , 357
p-form, 178
frame (of reference), 2, 192
free bodies (system of), 86
free-body diagram, 68-70
free index, 139
free particle, 81
friction force, 50
friction laws of Coulomb, 52
frictionless constraint, 55
functional, 362
fundamental formula of acceleration, 22
fundamental formula of rigid body kinematics,
19
fundamental principle of dynamics, 302, 303
fundamental tensor, 164

G
Galilean principle of inertia, 302
generalized coordinates, 72, 78-79, 340
generalized force, 106, 107, 357
generalized momentum, 359, 399
generalized trajectory, 73, 355, 372
generalized velocity, 73, 340, 355
generating function, 420, 427
geodesic, 212-213, 215, 371-374
given external forces, 36
given forces, 37
gradient of tensor field, 224-225
graduation, 188
Guldin (or Pappus) theorems, 271

H
hat, 180
Hamilton-Jacobi equation, 426, 430
Hamiltons variational principle, 369
Hamiltonian, 399
harmonic oscillator, 430
Hero of Alexandria, 360
hoist, 119

holonomic constraint, 76
holonomic system, 340
hyperstatic (collection of forces), 45
hypostatic (collection of forces)

I
ideal constraint, 96-97
ignorable coordinate, 360
impulse, 311
indefinite signature, 176
inertia cylinder, 280
inertia ellipse, 280
inertia ellipsoid, 279, 286-287
inertia ellipsoid equation, 278
inertia force, 342
inertia law, 302
inertial frame of reference, 304
integral (with respect to a mass distribution),
266
integral of f along a curve, 362
integral curve of field, 375
interaction forces, 37
internal forces of constraint, 37
internal mechanical action, 34
internal potential, 325
invariant Lagrangian, 376-379
invariant of dynam, 11
involutive function, 415
isostatic ( collection of foces), 45

J
Jacobis identity, 413, 414
Jacobis integral, 359, 404
Jacobis theorem, 428

K
kets, 137
kinematic dynam, 21
kinetic dynam, 307, 317
kinetic energy, 308, 326-328, 356
kinetic energy theorem, 323, 328-330
kinetic friction (coefficient of ), 53
kinetic friction force, 51
kinetic momentum, 307
Koenig expression of angular momentum, 308
Koenig expression of kinetic energy, 309, 324
Kronecker tensor, 153

458

L
Lagrange bracket, 415, 416
Lagranges equations, 347, 349
Lagrange multiplier, 351
Lagrangian, 349
Laplacian of function, 230
Laplacian of vector field, 231
laws of friction, 52
least time (principle of ), 360
Legendre transformation, 397, 399
length of arc, 211
Levi-Civita symbol, 216
Lie algebra, 413
line element, 201
line of action, 35
linear mapping, 135
linear momentum, 307
linear momentum theorem, 310
Liouvilles theorem, 407, 409
local reference frame, 304
local transformation, 375
lowering mapping, 166

M
mass of material system, 264
mass per unit area, 264
mass per unit length, 265
mass per unit volume, 264
material symmetry, 287
material system, 33
Maupertuis, principle, 361
mechanical action, 34
mechanical system, 33
method of isolation, 68
metric element, 201
metric tensor, 176
mixed tensor, 150, 152
modeled material system, 34
moment of dynam, 5
moment of inertia, 272-273
moment of inertia about origin, 273
moment of inertia about coordinate axes, 273
moment of inertia about coordinate planes, 273
moment of inertia about any axis, 277
moment of order k, 272
momentum canonically conjugate to q i , 359
motion, 300
motion law, 302
motion of mass center (theorem), 311
multilinear form, 306

Index
multiplication of dynam by scalar, 9
multiplication of dynams, 10
multiplication of a p-form by a scalar, 185
multiplication of a linear mapping by a scalar,
136
multiplication of a tensor by a scalar, 157

N
natural basis, 195
natural frame, 195
natural trajectory, 389
N-body problem, 393
neutral equilibrium, 110
Newtons first law, 302
Newtons second law, 302
Newtons third law, 303
nonholonomic system, 340
norm 176,
normal force, 49
(normal) friction force, 51
normal stress, 47
number of degrees of freedom of a constraint,
55
number of degrees of freedom of a system, 71
nutation angle, 87

O
one-form, 136
one-parameter group of diffeomorphisms, 375
orthogonal dynams, 10
orthogonal vectors, 165
orthonormal basis, 176

P
parallel forces, 44
parallel translation, 213
parallel transport, 213
parallelogram of forces (law of ), 303
partially antisymmetric tensor, 155
partially symmetric tensor, 154
phase point, 406
phase space, 403
phase spacetime, 409
pivot, 65, 68
pivoting moment, 49
plane support, 60, 68
point, 2

459

Index
point space, 2, 192
Poisson bracket, 409, 411
Poisson theorem, 414, 415
position vector, 3
postulate of initial conditions, 301
potential, 108
power of external forces, 323-330
power of internal forces, 323-330
precession angle, 87
pre-Euclidean space, 3
pre-Euclidean vector space, 164
primitive coordinates, 72
principal axes, 282, 286-287
principal axes of inertia, 283
principal moment of inertia, 283
principle of Galilean relativity, 305
principle of least action, 361
product of a dynam by a scalar, 8
product of a p-form by a scalar, 185
product of a tensor by a scalar, 157
product of dynams, 9
product of inertia, 272
proper motion (angle of ), 87
pseudo-Euclidean vector space, 176
pseudo-norm, 176
punctual constraint, 55

R
radius of gyration, 280
raising mapping, 167
reaction, 34
reciprocal basis, 169
reciprocal vector, 169
rectilinear constraint, 56
rectilinear coordinates, 194, 202
reduction of a vector system, 12
redundant constraint, 45
relative acceleration, 23
relative minimum, 362
relative velocity, 22
representation of a dynam, 6
representative point, 355
resultant of a dynam, 5
rheonomic constraint, 76, 341
Ricci identities, 206
Ricci theorem, 211
rigid body, 74
rigid body motion, 92
rolling moment, 49
rolling without slipping, 26
rotational kinetic energy, 328
Rouths method, 405-406
Routhian, 405

S
scalar, 158, 219
scalar invariant, 12
scalar multiplication, 164, 170
scalar product, 164, 172
scleronomic constraint, 76, 341
scleronomic system, 340
screw joint, 64, 68
second order contravariant tensor, 150
second order covariant tensor, 148
separable mechanical system, 432
separability (condition for ), 433
sharp mapping, 167
shear stress, 47
short path (principle of ), 360
signature, 176
skew-symmetric p-linear form, 178
slider-crank, 122
sliding dynam, 14, 16
sliding guide, 63, 68
sliding hinge, 61
sliding pivot,, 61, 68
sliding velocity, 26
smooth constraint, 55
space connected with a rigid body, 17
space of points and directions, 377, 389
spherical pendulum, 73
square of dynam, 10
stable equilibrium, 110
Stckels case of separability, 434-439
state space, 409
static friction (coefficient of ), 53
static friction force, 51
statically determinate structure, 45
Steiners theorem, 288-290
stiffness, 300
stress, 47
strict components, 183, 189, 216
subdivision, 269
sum of dynams, 7
sum of p-forms, 185
sum of tensors, 156
surface force, 49
symmetric tensor, 153-154
symplectic group, 425
symplectic matrix, 411
system of canonical equations (of Hamilton),
401
system of Eulers equations, 367
system of vectors, 3

460

Index

T
Tangent vector field, 374
Tangential force, 49
(tangential) friction force, 51
tension, 48
tensor algebra, 158
tensor criterion, 162-163
tensor density, 216
tensor field, 201
tensor inertia, 274
tensor multiplication, 157

(
tensor of type (
tensor of type (
tensor of type (
tensor of type (

tensor of type

tensor of type

0
1
1
0
0
2
2
0
1
1

) , 144
) , 145
) , 146
) , 149
) , 150

( ) , 152
q
p

tensor on point space, 199


tensor product, 143, 157
tensor product space, 147, 149, 151, 152
tensor space, 152
Torricelli theorem, 102
trace, 236
transformation of tensor components, 199
transport acceleration, 23
transport velocity, 22,87
transposed tensor, 151
translational kinetic energy, 328
trihedron of Copernicus, 304

uniform rectilinear translation, 306


unilateral constraint, 75
unstable equilibrium, 110

V
variational calculus, 368
variational derivative, 365, 371
vector 145, 219
q-vector, 188
vector invariant, 11
velocity, 214
velocity field, 19
velocity vector relative to 18, 19
virtual angular displacement, 93
virtual angular velocity vector, 93
virtual displacement, 33, 80
virtual kinematic dynam, 93
virtual rotation (vector of ), 85
virtual velocities (field of ), 85-86
virtual velocity, 33, 82, 84
virtual power, 94
virtual work, 94
virtual work principle, 98-102

W
wedge, 180
welded connection, 35
welded joint, 67

Z
U
unconstrained rigid body, 82
undetermined multipliers, 350-354

zero dynam, 7, 12
zero form, 184
zero tensor, 153

You might also like